Вы находитесь на странице: 1из 300

GEOLOGICAL SURVEY OF CANADA

BULLrnN 534

HOLOCENE CLIMATE AND ENVIRONMENTAL


CHANGE IN THE PALLISER TRIANGLE:
A GEOSCIENTIFIC CONTEXT FOR EVALUATING
THE IMPACTS OF CLIMATE CHANGE ON
THE SOUTHERN CANADIAN PRAIRIES
Edited by
D.S. Lemmen and R.E. Vanee

GEOLOGICAL SURVEY OF CANADA

BULLETIN 534

HOLOCENE CLIMATE AND ENVIRONMENTAL


CHANGE IN THE PALLISER TRIANGLE:
A GEOSCIENTIFIC CONTEXT FOR EVALUATING
THE IMPACTS OF CLIMATE CHANGE ON THE
SOUTHERN CANADIAN PRAIRIES

Edited by

D.S. Lemrnen and R.E. Vance

OHer Majesty the Queen in Right of Canada, 1999


Catalogue No. M42-534E
ISBN 0-660-17887-7
Available in Canada from
Geological Survey of Canada offices:
601 Booth Street
Ottawa, Ontario KIA OE8
3303-33rd Street N.W.
Calgary, Alberta T2L 2A7
101-605 Robson Street
Vancouver, B.C. V6B 553
A deposit copy of this publication is also available for reference
in selected public libraries across Canada

Price subject to change without notice

Cover illustration
The Red Deer River valley about 10 krn downstream from Drumheller,
Alberta. The river channel and terraces occupy a glacial meltwater
channel cut into Upper Cretaceous bedrock, with tributary streams having
incised deeply during the late Holocene. Agriculture is the dominant
landuse, and the potential impact of future climate changes on water and
soil resources is an issue of regional concern. Photograph by
D.J. Sauchyn. GSC 1999-049

Editors' addresses

D.S. Lemmen
Geological Survey of Canada
3303-33rd Street NW
Calgcity, Alberta
T2L 2A7
R.E. Vance
Natural Resources Canada
580 Booth Street
Ottawa, Ontario
K I A 0E4

Original nzanuscript submitted: 1998-08


Final version approved for publication: 1999-07

CONTENTS
Introduction
D.S. Lemmen and R.E. Vance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
An overview of the Palliser Triangle Global Change Project
D.S. Lemmen and R.E. Vance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
Geolimnology of the Great Plains of western Canada
W.M. Last . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .23
Groundwater in the Palliser Triangle: an overview of its vulnerability and
potential to archive climate information
V.H. Remenda and S.J. Birks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
Diatom-based salinity reconstructions from Palliser Triangle lakes: a summary of
two Saskatchewan case studies
S.E. Wilson and J.P. Smol . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . .67

Groundwater inputs to a closed-basin saline lake, Chappice Lake, Alberta


S.J. Birks and V.H. Remenda . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 1
Mineralogy, lithostratigraphy, and inferred geochemical history of North
Ingebrigt lake, Saskatchewan
Y.Shangand W.M.Last . . . . . . . . . . . . . . . . . . . . . . . . .
Late Holocene paleolimnology of Killarney Lake, Manitoba
K.-A. Richmond and L.G. Goldsborough . . . . . . . . .

. . . . . . . . . .95

. . . . . . . . . . . . . . . . 111

Multiproxy record of prairie lake response to climatic change and human activity,
Clearwater Lake, Saskatchewan
P.R. Leavitt, R.D. Vinebrooke, R.I. Hall, S.E. Wilson, J.P. Smol,
R.E. Vance, and W.M. Last. . . . . . . . . . . . . . . . . . . . . . . . . . . .
A postglacial plant macrofossil record of vegetation and climate change in
southern Saskatchewan
C.H. Yansa and J.F. Basinger . . . . . . . . . . . . . . . . . . . . . . .

. . . . . 125

...

The lithostratigraphic record of late Pleistocene-Holocene environmental change


at the Andrews site near Moose Jaw, Saskatchewan
A.E. Aitken, W.M. Last, and A.K. Burt . . . . . . . . . . . . . . . . . . . . . . . . . . 173
Sand dunes of the northern Great Plains of Canada and the United States
D.R.MuhsandS.A.Wolfe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183

Monitoring of dune activity in the Great Sand Hills region, Saskatchewan


S.A. Wolfe and D.S. Lemmen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
Using optical dating to determine when a sediment was last exposed to sunlight
D.J. Huntley and O.B. Lian. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .211
Activity cycle of parabolic dunes based on morphology and chronology from
Seward sand hills, Saskatchewan
P.P. David, S.A. Wolfe, D.J. Huntley, and D.S. Lemmen . . . . . . . . . .

. . . . . . . 223

Geomorphology of the western Cypress Hills: climate, process, stratigraphy, and theory
D.J. Sauchyn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
Groundwater influence on valley-head geomorphology, upper Battle Creek basin,
Alberta and Saskatchewan
C.D. Spence and D.J. Sauchyn . . . . . . . . . . . . . . . . . . . . . . . . . .

....

Origin and erosion of the Police Point landslide, Cypress Hills, Alberta
D.J. Sauchyn and H.L. Nelson . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .257
Geomorphic surfaces and postglacial landscape evolution of the Maple Creek
basin, Saskatchewan
W.J. Vreeken. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Author index

. . . . . . .267

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295

Introduction
Donald S. em men land Robert E. vance2

Lernrnen, D.S. and Vance, R.E., 1999: Introduction; in Holocene Climate and Environmental
Change in the Palliser Triangle: A Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lernrnen and R.E. Vance; Geological
Survey of Canada, Bulletin 534, p. 1-6.

The Palliser Triangle is the driest portion of the Canadian


Prairies, and one of the most climatically sensitive regions
in Canada. Extending from the southwestern corner of
Manitoba across all of southern Saskatchewan to
south-central Alberta, these lands were deemed by
Captain John Palliser to have marginal agricultural potential following his 1857-1860 expedition. Despite
Palliser's assessment, extensive Eurocanadian settlement
occurred during the late nineteenth and early twentieth
centuries, and it has since proven to be one of the world's
most important agricultural areas. Nonetheless, the social,
economic, and environmentalimpacts of severe drought in
the 1920s, 1930s, and 1980s highlighted the region's sensitivity to climatic variability.

Le Triangle de Palliser, qui est la partie la plus skche des Prairies


canadiennes, est une des regions du Canada dont le climat est le
plus sensible. S'Ctendant du coin sud-ouest du Manitoba
jusqu'au centre-sud de 17Alberta,en traversant tout le sud de la
Saskatchewan, ces terres Etaient consider6es par le capitaine
John Palliser, B I'issue de son expedition de 1857-1860, comme
&ant dotbes d'un potentiel agricole negligeable. En depit de
l'bvaluation de Palliser, une colonisation eurocanadienne intensive y a eu lieu au 19e et au dkbut du 20e sibcles et elles se sont
rCv616es depuis lors comme une des rkgions agricoles les plus
importantes du monde. Toutefois, les incidences sociales,
Bconorniques et environnementales des graves secheresses des
annkes 1920, 1930 et 1980 ont mis en lumikre la sensibilitk de
cette region h la variabilitk climatique.

As we enter the twenty-first century, concerns of


drought and associated impacts in the Prairie region
largely centre around the issue of global climate change.
Human activities are altering the composition of the
atmosphere, which in turn will likely affect climate. General Circulation Models predict that future global warming
will be most pronounced in northern regions and continental interiors, including the Great Plains of North America.
While uncertainties remain in these projections, particularly with respect to precipitation, there is general agreement that the Palliser Triangle is likely to become more
arid, and that drought frequency may increase. Evidence
of warming is already present in the regional climate
record.

A l'aube du 21e sibcle, les prkoccupations relatives aux


secheresses et 21 leurs incidences dans la rkgion des Prairies
s'articulent pour l'essentiel autour du problbme des
changements climatiques globaux. En effet, les activites
humaines modifient la composition de l'atrnosphkre, laquelle se
repercute 21 son tour selon toute vraisemblance sur le climat. Les
modbles de circulation gCnCrale predisent que le futur
rechauffement global sera le plus marque dans les regions
septentrionales et h I'intkrieur des continents, notamment la
region des grandes plaines d'AmCrique du Nord. Bien qu'il
subsiste des incertitudes dans ces prkdictions, en particulier eu
dgard aux precipitations, on s'accorde gCnCralement pour
admettre que le Triangle de Palliser deviendra probablement
plus aride et que la frkquence des ~Ccheressesy augmentera. Les
donnees climatiques rkgionales prksentent dbjh des indices de
rechauffement.

Change inevitably results in adaptation: natural systems will respond to climate change autonomously. However, human response has potential to be both strategic and
anticipatory if we can determine what the impacts of climate change are likely to be. Hence, to improve our

Le changement engendre inkvitablement I'adaptation. Les


systkmes naturels repondent aux changements climatiques de
manikre autonome, mais les reactions humaines pourront faire
preuve de strattgie et d'anticipation si nous sommes en mesure
de determiner les incidences probables des changements

' Terrain Sciences Division, Geological Survey of Canada, 3303-33rd Street NW, Calgary, Alberta T2L 2A7
Natural Resources Canada, 580 Booth Street, Ottawa, Ontario KIA OE4

GSC Bulletin 534

understanding of the role of climate, attention must be


focused on the dynamics of natural, social, and economic
systems. The study of analogues is a powerful tool for
understanding system dynamics: if we document the
impacts that resulted from past climate change and variability, we will have a much stronger basis for assessing
the impacts of future climate change. With respect to biophysical systems, the geological record contains unique
information on the impacts of a wide range of past climate
changes. It is this paleoenvironmental record, and the context it provides for assessing the impacts of future climate
change, that has been the focus of the Palliser Triangle
Global Change Project.

climatiques. Aussi, pour amCliorer notre comprChension du r61e


du climat, importe-t-il d'accorder une attention touteparticulibre
2 la dynamique des systbmes naturels, sociaux et Cconomiques.
Les analogies sont un outil puissant pour la comprChension de la
dynamique des systbmes; si nous recensons les incidences des
changements et de la variabilitC du climat dans le passC, nous
serons beaucoup mieux CquipCs pour Cvaluer les incidences des
changements climatiques h l'avenir. Pour ce qui est des
systkmes biophysiques, les donnCes gCologiques renferment des
informations uniques sur les incidences d'un vaste Cventail de
changements climatiques passCs. Ces donnCes palCoenvironnementales et le contexte qu'elles fournissent pour 1'Cvaluation des
changements climatiques futurs constituent l'axe central du
Projet sur le Triangle de Palliser et les changements climatiques
globaux.

This Bulletin contains 18 papers related to two major


objectives of the Palliser Triangle Project: 1) reconstmction of Holocene climatic and hydrological changes; and
2) evaluation of the relationships between climate and
landscape processes. Each paper was written to stand on
its own as a conwibution to a specific geoscience discipline. However, we hope that by bringing this set of papers
together in a regionally focused volume, we are able to
highlight the need for interdisciplinary and multidisciplinary studies when addressing global change
issues. The papers in this volume alone do not form a comprehensive summary of project research. Many results,
particularly related to the first objective above, have been
previously published. Readers are directed to these additional publications through references provided in this
volume.

Ce bulletin contient 18 textes portant sur deux des principaux


objectifs du Projet sur le Triangle de Palliser : (1) la reconstitution des changements climatiques et hydrologiques d e
1'Holocbne et (2) 1'Cvaluation des rapports entre le climat et les
processus fa~onnantle paysage. Chaque texte a Ct6 rCdigC de
manibe 2 fournir une contribution autonome B une discipline
gkoscientifique spkcifique. Nous espCrons qu'en rassemblant
cet ensemble de textes en un volume B orientation regionale nous
serons en mesure de mettre en relief la nCcessitC d'effectuer des
Ctudes interdisciplinaires et multidisciplinaires pour aborder les
problbmes relatifs aux changements climatiques globaux. 11faut
Cgalement noter que les textes de ce seul volume ne constituent
pas un rCsumC complet des recherches menCes dans le cadre du
projet. De nombreux rCsultats relatifs en particulier au premier
objectif CnumCrC plus haut ont kt6 publies antkrieurement. Ces
publications supplCmentaires sont signalCes au lecteur dans les
rCfCrences incluses dans ce volume.

The first paper, by Lemmen and Vance (1999), is an


overview of the project and a broad synthesis of major
results. The subsequent papers provide thematic reviews
of environmental settings, systems, and methodologies
critical to the understanding of past regional climate
change, as well as results of detailed, site-specific studies
supported by the project.

Le premier texte, de Lernrnen et Vance (1999), est un rCsumC


du projet et une vaste synthkse des principaux rksultats. Les
textes suivants prksentent des rCsumCs thCmatiques des cadres
environnementaux et des systbmes et mCthodologies qui sont
d'importance critique pour la comprChension des changements
climatiques rCgionaux passCs, ainsi que les rCsultats d'6tudes
dCtaillCes sur des sites spCcifiques soutenues par le projet.

The nine papers that follow the project overview,


which make up the first half of the volume, are devoted to
records of past climate and hydrological change documented in paleolimnological and hydrogeological studies.
Introductory papers familiarize the reader with the physical lake environments of the Canadian Prairies (Last,
1999) and the critical role that groundwater plays in maintaining these lakes (Remenda and Birks, 1999). These
papers detail the diversity and complexity of hydrological
systems, highlighting difficulties involved in determining
a coherent paleoclimatic signal from lake sediments. They
emphasize the need for a thorough understanding of the
hydrological dynamics when formulating paleoenvironmental interpretations. An introduction to biological
proxy data is provided by Wilson and Smol (1999), who
summarize diatom studies from two lakes within the
Palliser Triangle. The similarities and differences evident
in these records underscore the importance of examining
sites from a variety of topographic, climatic, and

Les neuf textes qui suivent le rCsumC de projet et qui constituent donc la premibre moitiC de ce volume sont consacrCs aux
donnCes sur les changements climatiques et hydrologiques
passks qui sont dCcrites dans les Ctudes palColimnologiques et
hydrogCologiques. Des textes introductifs familiarisent le
lecteur avec le milieu physique des lacs des prairies (Last, 1999)
et avec le r81e critique que jouent les eaux souterraines pour la
persistance de ces lacs (Remenda et Birks, 1999). Ces textes
expliquent en dCtail la diversit6 et la complexitC des systbmes
hydrologiques, tout en mettant en relief les difficult& que
comporte la dktermination d'un signal palCoclimatique cohkrent
h partir de sidirnents lacustres. 11s souligent la nkcessitC d'une
profonde comprChension d e la dynamique hydrologique
lorsqu'il s'agit de formuler des interprCtations environnementales. Une introduction aux donnkes indirectes en biologie
est fournie par Wilson et Smol (1999) qui rCsument des Ctudes
sur les diatomees dans deux lacs du Triangle de Palliser. Les
similitudes et diffkrences qui sont manifestes dans ces donnCes
soulignent 2quel point il est important d'examiner des sites dans
des cadres topographiques, climatiques et hydrologiques

D.S. Lemmen and R.E. Vance

hydrological settings in order to differentiate variability


related to regional factors, including climate, from local
dynamics, such as groundwater discharge.

diffkrents afin de pouvoir distinguer la variabilitC like des


facteurs rCgionaux, y compris le climat, de la dynarnique locale,
par exemple des Ccoulements d'eau souterraine.

The next two papers present results of detailed hydrogeological and minerological studies. Despite the important role that groundwater plays in controlling lakes in this
environment, only Chappice Lake has a groundwater
monitoring study (Bisks and Remenda, 1999) to support a
detailed paleoenvironmental record. While the duration
and scope of this groundwater study precluded the development of a quantitative hydrogeological budget, the
results emphasize the importance of such data for interpreting paleohydrological conditions. Similar studies are
required elsewhere on the northern Great Plains if a
regional paleoclimatic signal is to be derived.

Les deux textes suivants prCsentent les rksultats d'Ctudes


hydrogtologiques et mintralogiques dCtaillCes. En dtpit du r6le
important que jouent les eaux souterraines dans la regulation des
lacs de ce milieu, seul le lac Chappice a fait l'objet d'une Ctude de
contr6le des eaux souterraines (Birks et Remenda, 1999)
permettant de recueillir des donnCes environnementales
dCtaillCes. MEme si la dur6e et l'ampleur de cette Ctude sur les
eaux souterraines ont emp$chC l'klaboration d'un budget
hydrogCologique quantitatif, les rCsultats soulignent l'importance de telles donnCes pour l'interprktation des conditions
paltohydrologiques. ~ e sCtudes
semblables sont ntcessaires
ailleurs dans les grandes plaines septentrionales si l'on veut
dCtecter un signal palCoclimatique regional.

Groundwater is also a key factor influencing the stratigraphic record of North Ingebrigt lake (Shang and Last,
1999), one of many hypersaline lakes and playas found on
the northern Great Plains (see overview paper by Last,
1999). Such basins have traditionally received limited
attention in paleoenvironmental research, as the high
salinity restricts or even precludes most organisms used as
proxy climate indicators. However, this paper exemplifies
how the salts, which comprise the majority of the sediment
record, provide detailed paleolimnological and
paleohydsological data (particularly related to brine composition), and valuable paleoclimatic insights (in this case
reconstruction of relative humidity).

Les eaux souterraines constituent un facteur clC influenpnt


Cgalement les donnCes stratigraphiques du lac North Ingebrigt
(Shang et Last, 1999), un des nombreux lacs hypersalins/playas
que l'on trouve dans les grandes plaines septentrionales (voir
l'aperqu par Last, 1999). Dans le pass6, on a gCnCralementpr$tC
peu d'attention h de tels bassins en recherche pal6oenvironnementale, car la salinitC ClevCe restreint, voire ernpeche la
prCsence de la plupart des organismes utilisCs comme
indicateurs indirects du climat. Ce texte explique toutefois
pourquoi les sels, qui constituent la plus grande partie des dCp6ts
~Cdimentaires,fournissent des donnCes palColimnologiques et
palCohydrologiques dttailltes (likes en particulier 9 la composition des saumures) et d'utiles aperps palCoclimatiques (dans ce
cas, la reconstitution de l'humiditk relative).

Biological proxy data are emphasized in the following


two papers. The late Holocene sedimentary record in
Killarnev Lake in southeastern Manitoba documents
changes in diatoms and pigments that are correlative with
sites to the west, suggesting that there is a recognizable
regional climate signal (Richmond and Goldsborough,
1999).In addition, this paper provides a context for assessing the impact of recent human activities in the region, particularly in terms of water quality. The study of Clearwater
Lake by Leavitt et al. (1999) is an example of the insights
that can be gained through collaborative research using
multiple proxy indicators. Through integration of diatom,
pigment, plant macrofossil, and mineralogical data, the
study provides more details of past changes than would be
evident from any single indicator. They demonstrate that
changes in land use over the past century dominate the climatesignal at this site. This highlights arecurrent theme of
the Palliser Project: that a fundamental requirement of any
rigorous paleoclimatic reconstruction must be an objective assessment of all factors in environmental change,
rather than simply the role of climate.

Les donnCes biologiques indirectes sont mises en relief dans


les deux textes suivantsl Le lac Killarney, dans le sud-est du
Manitoba, qui renferme des dCp6ts de 1'Holocbne supkrieur,
ttmoigne de changements relatifs aux diatomties et aux pigments
qui sont en corrClation avec des sites 9 l'ouest, ce qui laisse
penser qu'il existe un signal climatique rkgional reconnaissable
(Richmond et Goldsborough, 1999). Cet article fournit en outre
un contexte permettant d'Cvaluer les incidences des activitCs
humaines rCc&tes dans la rCgion, en particulier relativement la
qualit6 de l'eau. L'Ctude de Leavitt et al. (1999) sur le lac
Clearwater fournit un exemple des a p e r p s qu'offrent les
recherches concertCes utilisant des indicateurs indirects multiples. Gr2ce 9 1'intCgration de donntes sur les diatomCes, les
pigments et les macrofossiles vCgCtaux et de donnCes min6alogiques, cette Ctude fournit plus de dCtails sur les changements
passCs que ne le rkvtlerait un indicateur unique. Leavitt et al.
(1999) montrent que les changements dans l'utilisation du sol
depuis un sikcle dominent le signal climatique de ce site. Cela
illustre une thCmatique rCcurrente du projet sur le Triangle de
Palliser, B savoir que 1'Cvaluation objective de tous les facteurs
de changement environnemental, par opposition 9 la seule prise
en consid6ration du r61e du climat, doit constituer une
exigence fondamentale de toute reconstitution palCoclimatique
ngoureuse.

The final two papers of the first set utilize data from an
excavated prairie pothole to derive insights about environmental conditions during the latest Pleistocene-early
Holocene. The extremely diverse plant macrofossil
assemblages at this site (Yansa and Basinger, 1999)

Les deux derniers textes de cette sCrie utilisent des donntes


provenant d'une cuvette des Prairies excavCe afin de trouver des
indices sur les conditions environnementales pendant le
PlCistochne sommital et 1'Holocbne inftirieur. L'extreme
diversit6 des assemblages de macrofossiles vCgCtaux de ce site

GSC Bulletin 534

document, among other things, the postglacial migration


of white spruce through southwestern Saskatchewan. The
roles of climate, hydrology, and landscape instability are
all considered in the reconstruction of local conditions
extending from deglaciation to about 6000 years ago. The
lithostratigraphy of the same site, as interpreted by Aitken
et al. (1999), provides an interesting contrast. While strong
similarities between the studies exist, mineralogical data
suggest somewhat different interpretations of climate
dynamics than those derived from the plant macrofossil
data. Again, these results emphasize the importance of
integrating multiple proxy indicators to achieve realistic
paleoenvironmental interpretations.

(Yansa et Basinger, 1999) permet notamment de reconstituer la


migration postglaciaire de 1'Cpinette blanche dans le sud-ouest
de la Saskatchewan. On a tenu compte des r6les respectifs du
climat, de l'hydrologie et de l'instabilit6 des paysages pour
reconstituer les conditions locales 8 partir de la dCglaciation
jusqu78il y a environ 6 000 ans. La lithostratigraphie du meme
site, telle qu'interpretke par Aitken et al. (1999), fournit un
contraste inttressant. Bien qu'il existe de fortes similitudes
entre les deux Ctudes, les donnCes minCralogiques suggbrent des
interprktations quelque peu differentes de la dynamique
climatique que celles suggCrCes par les donnCes macrofossiles
vCgCtales. La encore, ces rCsultats soulignent l'importance de
I'intCgration de multiples indicateurs indirects pour parvenir 2
des interprktations palCoenvironnementales rtalistes.

The papers in the second half of the volume focus on


geomorphic processes in the Palliser Triangle, and how
these respond to climate forcing. An overview of regional
eolian, fluvial, and mass-wasting systems, as well as soil
redistribution, was presented by Lemmen et al. (1998),
based on literature predating the Palliser Project. That
review concluded that eolian landscapes are the most sensitive to climatic variability, and that much of the region
lies near a threshold of extensive eolian activity.

Les textes de la seconde moiti6 de ce volume s'articulent


autour des processus gComorphologiques dans le Triangle de
Palliser et de la manihe dont ces processus repondent au for~age
climatique. Une vue d'ensemble des systbmes Coliens, fluviaux
et de mouvements de masse rCgionaux ainsi que de la rtpartition
des sols a CtC pr6sentCe par Lemmen et al. (1998), sur la base
d'une documentation antkrieure au projet sur le Triangle de
Palliser. Les auteurs concluent que les paysages Coliens sont les
plus sensibles h la variabilitk climatique et qu'une bonne partie
de la region est proche du seuil d'une activitC Colienne intense.

The first four papers in this set focus on eolian environments. Muhs and Wolfe (1999) place sand dunes of the
Palliser Triangle within the broader context of the northern Great Plains. They review the major controls on dune
formation and activity, evidence of past dune activity from
geological, pedological, and historical sources, and finally
the value of dune areas in developing paleoenvironmental
reconstructions. Among many important conclusions is
that most dune fields on the northern Great Plains are not
simply relics of Late Pleistocene deglaciation, but have
been active during the late Holocene and are only marginally stable at present. A second overview paper with a
strong eolian theme concerns optical dating (Huntley and
Lian, 1999).This comparatively new dating technique has
provided a breakthrough in our understanding of sand
dune activity (see paper by David et al., 1999). By providing an age of the last time sediment was exposed to sunlight, the technique is able to directly document periods of
past eolian activity, in contrast to other techniques, such as
radiocarbon dating of paleosols, which document periods
of minimal eolian activity.

Quatre textes de ce bulletin sont axes sur les milieux Coliens.


Muhs et Wolfe (1999) replacent les dunes de sable du Triangle
de Palliser dans le contexte plus vaste des grandes plaines
septentrionales. 11s examinent les contr6les principaux sur la
formation et I7activitCdes dunes, les indices d'activitt des dunes
dans le pass6 h partir de sources g6010giques, pCdologiques et
historiques et, enfin, l'utilitt des rCgions de dunes pour les
reconstitutions palCoenvironnementales. Une des nombreuses
conclusions importantes qu'ils tirent est que la plupart des
champs de dunes des grandes plaines septentrionales ne sont pas
simplement des vestiges de la deglaciation du PlCistocbne
supCrieur, mais ont CtC actifs pendant 1'Holocbne supCrieur et
sont h peine stables aujourd'hui. Un deuxibme texte servant de
tour d'horizon et B forte thematique Colienne concerne la
datation optique (Huntley et Lian, 1999). Cette technique de
datation relativement rCcente a fourni une percCe dans notre
comprChensionde I'activitC des dunes de sable (voir David et al.,
1999). En permettant de dater la dernibre exposition d'un
sediment h la lumibre solaire, cette technique peut renseigner
directement sur les pCriodes d'activit6 6olienne du passC, ce qui
contraste avec d'autres techniques telles que la datation au
radiocarbone des palCosols, qui fournissent des renseignements
sur des pCriodes d'activit6 Colienne minimale.

The other two papers on eolian environments summarize recent field studies in the Palliser Triangle. The present level of sand dune activity in the central Palliser
Triangle lends itself well to monitoring studies (Wolfe and
Lemmen, 1999), providing baseline data that is critical to
evaluating both paleoenvironmental interpretations and
the controls on modern eolian activity. Despite the limited
duration of the data set, insights are provided into the seasonality of dune activity, as well as rates of blowout erosion and slipface advance. Morphological and
geochronological (optical dating) data are used by David
et al. (1999) to develop the concept of an 'activity cycle'

Les deux textes ultCrieurs sur les milieux Coliens rCsument


des Ctudes de terrain rkcentes dans le Triangle de Palliser. Le
niveau actuel d'activitk des dunes de sable dans la partie centrale
du Triangle de Palliser se prete bien aux Ctudes de contr6le
(Wolfe et Lemmen, 1999), car il permet d'obtenir des donnCes
de base qui sont d'importance critique pour 1'Cvaluation des
interprktations palCoenvironnementales et des contr6les de
I'activitC Colienne contemporaine. En dCpit du caractkre limit6
dans le temps de l'ensemble de donnees, les auteurs fournissent
des aperys sur la saisonnalitC de l'activitk des dunes ainsi que
sur les taux d'Crosion des cuvettes de deflation et d'avande
des talus croulants. Des donnees morphologiques et

D.S. Lemmen and R.E. Vance

for parabolic dunes. Moisture availability, reflected by


fluctuations in the groundwater table, is considered the
ultimate control on dune activity. The conceptual model
presented in this paper incorporates fundamental
geomorphic principals such as threshold response and
responselrelaxation times that are required for assessing
the potential response of dune systems to future climate
change.

gCochronologiques (datation optique) sont utilisCes par David


et al. (1999) pour Claborer le concept de <<cycled'activitb pour
les dunes paraboliques. La disponibilit6 de l'humiditC, dont
tCmoignent des fluctuations d e la nappe phrkatique, est
considCrCe comme le contr8le ultime sur 1'activitC des dunes. Le
modble conceptuel prCsentC dans cet article incorpore des
notions gComorphologiques fondamentales telles que la rCaction
au seuil ainsi que le temps de rCaction et la pCriode de relaxation,
1'Cvaluation
i
de la rCaction
notions qui sont nkcessaires ?
potentielle des systbmes dunaires aux futurs changements
climatiques.

The next three papers focus on geomorphic data from


the Cypress Hills. The first of these (Sauchyn, 1999)
places previous observations within a broader framework
of theoretical geomorphology. In discussing the importance of scale (both spatial and temporal) and the bias of
the stratigraphic record to low-frequency, high-magnitude
(e.g. extreme) events, this paper highlights the challenge
facing the geomorphologist when attempting to address
the impacts of climate on landscape processes. The subsequent two papers underscore Sauchyn's views while
examining controls on fluvial and mass-wasting processes
in the Cypress Hills. Spence and Sauchyn (1999) integrate
field monitoring with geographic information system
(GIs) analysis to derive insights concerning the role of
groundwater discharge on local geomorphic processes.
Ongoing monitoring studies are the focus of Sauchyn and
Nelson's (1999) paper on Police Point landslide, the largest historic landslide in the Cypress Hills. This work documents significant 'downstream' impacts of the slide more
than tw; decades after the initial failure occurred. exemplifying the significant response times o f many
geomorphic systems. This fundamental point must be considered when addressing future adaptation needs.

Les trois textes suivants s'articulent autour des donnCes


gComorphologiques sur les collines Cypress. Le premier replace
les observations antkrieures dans un cadre plus vaste de
gComorphologie thCorique (Sauchyn, 1999). Soulignant
l'importance des Cchelles utilisCes (spatiale et temporelle) et de
la polarisation des donnCes stratigraphiques vers les CvCnements
de basse frCquencelforte magnitude (c'est-A-dire les CvCnements
extrEmes), cet article met en tvidence le dCfi qui se prCsente au
gComorphologue lorsqu'il aborde 1'Ctude des incidences du
climat sur les processus faqonnant le paysage. Les deux textes
suivants prksentent un approfondissement des id6es de Sauchyn
ainsi qu'un examen des contrales des processus fluviaux et des
mouvements en masse dans les collines Cypress. Spence et
Sauchyn (1999) intbgrent le contrale sur le terrain et l'analyse au
moyen du systbme d'information gkographique (SIG) afin
d'Clucider le r81e de 1'Ccoulement des eaux souterraines sur les
processus gCo~norphologiqueslocaux. Les Ctudes de contrale en
cours constituent le sujet principal abordC par l'article de
Sauchyn et Nelson (1999) sur le glissement de la pointePolice, le
plus important glissement d e terrain pendant la pCriode
historique dans les collines Cypress. Les auteurs y dCcrivent les
incidences importantes << en aval >> du glissement plus de deux
dCcennies aprbs la rupture initiale, ce qui illustre la longueur des
temps de rCaction de nombreux systbmes gComorphologiques. I1
importe de tenir compte de ce point essentiel pour 1'Cvaluation
des futurs besoins d'adaptation.

The final paper of the volume (Vreeken, 1999) is a


comprehensive evaluation of processes and events controlling Holocene landscape evolution in a closed drainage
basin. Focused primarily on fluvial systems, this study
demonstrates that, for this basin, geological and
geomorphic controls dominate climate at a temporal scale
of centuries to millennia. It also shows that simple temporal correlation between climatic and geomorphic events
does not demonstrate a causal relationship. With respect to
impacts of future climate change, the paper emphasizes
that only through knowledge of all factors controlling the
system of interest can the specific role of climate be
assessed.

Le texte final de ce volume (Vreeken, 1999) est un examen


dCtaillC des processus et CvCnements rCgissant 1'Cvolution des
paysages holocbnes dans un bassin versant ferm6. AxCe
essentiellement sur les systbmes fluviaux, cette Ctude montre
que dans le cas de ce bassin, les contrales gCologiques et
gComorphologiques dorninent le climat B 1'6chelle temporelle
des sibcles ou des mill6naires. Elle montre Cgalement que la simple corrClation temporelle entre CvCnements climatiques et
gComorphologiques ne permet pas d'Ctablir l'existence de
relations causales. En ce qui concerne les incidences des futurs
changements climatiques, c e texte souligne que seule la
connaissance de tous les facteurs influant sur le systbme CtudiC
peut permettre 1'Cvaluation du r61e spCcifique du climat.

As a final note, readers should be aware that review


manuscripts for some contributions were submitted as
early as March 1997. Hence some aspects of the work presented here may have been expanded upon or superceded
by additional research prior to release of this volume.
Readers are asked to contact individual contributors concerning possible follow-up to material presented herein.

Enfin, le lecteur doit Etre conscient du fait que, pour certaines


contributions, les manuscrits ont CtC soumis pour rCvision dbs
mars 1997. Aussi certains aspects des travaux prCsentCs ici
ont-ils pu Ctre dCveloppCs ou dCpassCs par des recherches
ultkrieures avant la publication de ce volume. Nous recommandons au lecteur de communiquer avec les collaborateurs
individuels pour se renseigner au sujet du suivi des travaux
prCsentCs ici.

GSC Bulletin 534

Aitken, A.E., Last, W.M., and Burt, A.K.


1999: The lithostratigraphic record of late Pleistocene-Holocene environmental change at the Andrews site near Moose Jaw, southern
Saskatchewan;in Holocene Climate and Environmental Change in
the Palliser Triangle: A Geoscientific Context for Evaluating the
Impacts of Climate Change on the Southern Canadian Prairies, (ed.)
D.S. Le~nmenand R.E. Vance; Geological Survey of Canada, Bulletin 534.
irks, S.J. and Remenda, V.H.
1999: Groundwater inputs to a closed-basin saline lake, Chappice Lake,
Alberta; in Holocene Climate and Environmental Change in the
Palliser Triande: A Geoscientific Context for Evaluating the
Impacts of ~ l i m a t Change
c
on the SouthernCanadian ~rairiesr(ed.)
D.S. Lemmen and R.E. Vance; Geological
- Survev of Canada, Bulletin 534.
David, P.P., Wolfe, S.A., Huntley, D.J., and Lemmen, D.S.
1999: Activity cycle of parabolic dunes based on morphology and chronology~of-Sewardsand hills, Saskatchewan; in Holocene Climate
and Environmental Change in the PalliserTriangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.
Huntley, D.J. and Lian, O.B.
1999: Determining when a sediment was last exposed to sunlight using
optical dating; in Holocene Climate and Environmental Change in
the Palliser Triangle: A Geoscientific Context for Evaluating the
Impacts of Climate Changeon the Southern Canadian Prairies, (ed.)
D.S. Lemrnen and R.E. Vance; Geological Survey of Canada, Bulletin 534.
Last, W.M.
1999: Geolirnnology of the Great Plains of western Canada; in Holocene
Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Leavitt, P.R., Vinebrooke, R.D., Hall, R.I., Wilson, S.E., Smol, J.P.,
Vance, R.E., and Last, W.M.
1999: Multiproxy record of prairie lake response to climate change and
human activity, Clearwater Lake, Saskatchewan; in Holocene Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemrnen and
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Lemmen, D.S. and Vance, R.E.
1999: An overview of the Palliser Triangle Global Change Project; in
Holocene Climate and Environmental Change in the Palliser Triangle: A GeoscientificContext for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Lemmen, D.S., Vance, R.E., Campbell, I.A., David, P.P.,
Pennock, D.J., Sauchyn, D.J., and Wolfe, S.A.
1998: Geomorphic systems of the Palliser Triangle, southern Canadian
Prairies: description and response to changing climate; Geological
Survey of Canada, Bulletin 521,72 p.
Muhs, D.R. and Wolfe, S.A.
1999: Sand dunes of the northern Great Plains of Canada and the United
States; in Holocene Climate and Environmental Change in the
Palliser Triangle: A Geoscientific Context for Evaluating the
Impacts of Climate Change on the Southern Canadian Prairies, (ed.)
D.S. Lemmen and R.E. Vance; Geological Survey of Canada, Bulletin 534.
Reinenda, V.H. and Birks, S.J.
1999: Groundwater in the Palliser Triangle: an overview of its vulnerability and potential to archive climate information; in Holocene CLimate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and
R.E. Vance; Geological Survey of Canada, Bulletin 534.

Richmond, K.-A. and Goldsborough, L.G.


1999: Late Holocene paleolimnology of Killarney Lake, Manitoba; in
Holocene Climate and Environmental Change in tlie Palliser Triangle: A GeoscientificContext for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lem~nenand
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Sauchyn, D.J.
1999: Geomorphology of the western Cypress Hills: climate, process,
stratigraphy and theory; in Holocene Climate and Environmental
Change in the Palliser Triangle: A GeoscientificContext for Evaluating the Impacts of Climate Change on the Southern Canadian
Prairies, (ed.) D.S. Lem~nenand R.E. Vance; Geological Survey of
Canada, Bulletin 534.
Sauchyn, D.J. and Nelson, H.L.
1999: Origin and erosion of the Police Point landslide, Cypress Hills,
southeastern Alberta; in Holocene Climate and E~ivironme~ital
Change in the Palliser Triangle: A GeoscientificContext for Evaluating the Impacts of Climate Change on the Southern Canadian
Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological Survey of
Canada, Bulletin 534.
Shang, Y. and Last, W.M.
1999: Mineralogy, lithostratigraphy,and inferred geochemical history of
North Ingebrigt lake, Saskatchewan; in Holocene Climate and
Environmental Change in the Palliser Triangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemrnen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.
Spence, C.D. and Sauchyn, D.J.
1999: Groundwater influence on valley-head geomorphology,upper Battle Creek basin, Alberta and Saskatchewan; in Holocene Climate
and Environmental Change in thepalliser Triangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Le~nmenand R.E. Vance;
Geological Survey of Canada, Bulletin 534.
Vreeken, W.J.
1999: Geomorphic surfaces and postglacial landscape evolution of the
Maple Creek basin, Saskatchewan; in Holocene Climate and Environmental Change in the Palliser Triangle: A GeoscientificContext
for Evaluating the Impacts of Climate Change on the Southern
Canadian Prair~es,(ed.) D.S. Lemmen and R.E. Vance; Geological
Survey of Canada, Bulletin 534.
Wilson, S.E. and Smol, J.P.
1999: Diatom-based salinity reconstructionsfrom Palliser Triangle lakes:
a summary of two Saskatchewan case studies; in Holocene Climate
and EnvironmentalChange in the Palliser Triangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.
Wolfe, S.A. and Lemmen, D.S.
1999: Monitoring of sand dune activity in the Great Sand Hills region,
southwestern Saskatchewan; in Holocene Climate and Environmental Change in the Palliser Triangle: A GeoscientificContext for
Evaluating the Impacts of Climate Change on the Southern
Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological
Survey of Canada, Bulletin 534.
Yansa, C.H. and Basinger, J.F.
1999: A postglacial plant macrofossil record of vegetation and climate
change in southern Saskatchewan; in Holocene Climate and Environmental Change in the Palliser Triangle: A GeoscientificContext
for Evaluating the Impacts of Climate Change on the Southern
Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological
Survey of Canada, Bulletin 534.

A11 overview of the Palliser Triangle


Global Change Project
Donald S.

em men' and Robert E. vance2

Lemmen, D.S. and Vance, R.E., 1999: An overview of the Palliser Triangle Global Change
Project; in Holocene Climate and Environmental Change in the Palliser Triangle: A
Geoscientijic Contextfor Evaluating the Impacts of Climate Change on the Southern Canadian
Prairies, (ed.)D.S. Lemrnen and R.E. Vance; Geological Survey of Canada, Bulletin 534,p. 7-22.

Abstract: The Palliser Triangle is the driest portion of the Canadian Prairies, and one of the most climatically sensitive regions in Canada. The potential biophysical impacts of future climate changes are addressed
in the Palliser Triangle Global Change Project through an improved understanding of Holocene climate and
hydrological changes, and associated landscape response.
Water availability, particularly in relation to groundwater fluctuations, is the single most important
factor controlling regional environmental change. Three major intervals are defined for the Holocene,
reflecting climate, hydrological, and human factors, and geomorphic activity during those intervals is
assessed. Projections of future geological impacts must be based on a thorough understanding of
hydrological and geomorphic system dynamics, including the importance of thresholds and antecedent
conditions. Systems that are still responding to past major disturbances are unlikely to show a predictable
response to climate change. Significant response and relaxation times in many systems indicate that some
climate impacts will not be immediately apparent, but may have considerable long-term consequences.

RCsumC : Le Triangle de Palliser constitue la partie la plus skche des Prairies canadiennes et une des
r6gions canadiennes dont le climat est le plus sensible. L'Ctude des incidences biophysiques potentielles des
futurs changements climatiques est abordCe par le Projet sur le Triangle de Palliser et les changements
climatiques globaux, qui vise i amtliorer notre comprehension du climat et des changements hydrologiques
au cours de 1'Holockne et des rtactions associees du paysage.
La disponibilitCde l'eau en gCnCra1, et les fluctuations des eaux soutei-raines en particulier, constituent le
facteur de contr8le des changements environnementaux rkgionaux le plus important. Trois intervalles
principaux, refletant les facteurs climatologiques, hydrologiques et humains, ont Ctt dtfinis dans
17Holocbne;l'activitk gComorphologique au cours de ces intervalles a Cgalement t t t CtudiCe. Les
projections des incidences gCologiquesfutures doivent &trefondtes sur une comprehension approfondie de
la dynamique des systkmes hydrologiques et gComorphologiques, notamment de l'importance des seuils et
des conditions anterieures. U est peu probable que les systkmes qui rkagissent encore aux perturbations
majeures du pass6 manifestent une reaction prkvisible aux changements climatiques. La longueur des
temps de reaction et des pCriodes de relaxation de nombreux systkmes indique que certaines incidences
climatiques ne seront pas immediatement manifestes, bien qu'elles puissent avoir des consCquences i long
terme considkrables.

' Terrain Sciences Division, Geological Survey of Canada, 3303-33rd Street NW, Calgary, Alberta T2L 2A7
Natural Resources Canada, 580 Booth Street, Ottawa, Ontario KIA 0E4

GSC Bulletin 534

INTRODUCTION
Extending from southwestern Manitoba to south-central
Alberta, the Palliser Triangle is the driest part of the Canadian
Prairies (Fig. 1). The following assessment of the region was
provided by Captain John Palliser following his land survey
of 1857-1 860 (Spry, 1968):

. . .in the central part of the continent there is a region, desert,


or semi-desert in character, which can never be expected to

become occupied by settlers ....Although there are fertile


spots throughout its extent it can never be of much advantage to us as a possession.

For the most part, the twentieth century has proven


Palliser to have been overly pessimistic, as the region that
bears his name forms much of the highly productive
Canadian wheat belt. Nonetheless, the social, economic, and
environmental impacts of historical drought, exemplified by
the 1920s, 1930s, and 1980s, highlight the region's sensitivity to climatic variability (Wheaton et al., 1992).
Concerns over the impacts of climate variability on
human activities are amplified when placed in the context of
projected future climate change. There is general agreement
among General Circulation Model (GCM) simulations that
enhanced atmospheric concentrations of greenhouse gases
will be associated with increased temperatures and decreased
summertime precipitation across much of the North
American Great Plains (e.g. Karl et al., 1991). This, in turn,
could cause increased frequency and magnitude of drought,
by far the greatest climatic hazard to economic sustainability

on the Canadian Prairies (Herrington et al., 1997; Nemanishen,


1998). While major advances have been made in climate
modelling over the last decade, confidence in some projections, particularly those of precipitation and soil moisture,
remains comparatively low. Furthermore, regional climate
models (RCMs), which can provide data at the spatial scale
necessary to address the regional impacts of climate change,
remain in the early stages of development (Laprise et al.,
1997).
A complementary approach to the modelling of future climates and associated impacts is the study of past environmental change. Paleoenvironmental research provides numerous
unique contributions to our understanding of climate change.
In the case of climate dynamics, a paleoenvironmental perspective spanning several millennia is required to define the
full range of climate variability, and hence provide context
for recent climate trends (e.g. Lemrnen et al., 1997). Furthermore, paleoclimatic data is critical for differentiating
between anthropogenic and other (natural) drivers of climate
change (Mann et al., 1998), as well as providing a basis for
validation of GCMs through hindcasting (Gajeweski et al., in
press).
In addition to providing proxy climate data, paleoenvironmental research provides information about the
response of biophysical systems to climate forcing. In documenting system response (autonomous adaptation (Smit
et al., in press)) to a range of past climatic changes, these studies provide a basis for assessing climatic sensitivity. Given
that the paleoenvironmental record encompasses a greater
range of climatic variability than documented in the historic

A A ~ * *

Mtn.

,"d

*a.

Figure 1. The Palliser Triangle (bold dashed line) as defined by John Palliserduring the 1857-1860 British
North American Exploring Expedition (from Spry, 1968). Isolines denote ratio of average annual
precipitation to potential transpiration, based on 30-year means (climate data from Environment Canada,
1993). Lower values indicate greater aridity; values have been interpolated for the Cypress Hills. Shaded
area is the Brown Chernozemic Soil Zone; bold dotted line is the continental drainage divide.

D.S. Lemmen and R.E. Vance

(instrumental) climate record, it potentially contains analogues of projected future climates. The great advantage of
such information, compared to model-based analyses, is that
it is based upon what has actually happened, as opposed to
projections of what may happen. Furthermore, the
paleoenvironmental record will reflect the integrated
response of a complex, dynamic system that may not be adequately represented in static models. An improved understanding of past climate changes and associated
environmental response therefore provides critical context
for any assessment of the impacts of future climate change.

THE PROJECT
The Palliser Triangle Integrated Research and Monitoring
Area (IRMA) was established in recognition of the climatic
sensitivity of the Prairies, projections of significant future climate change in that region, and the role that
paleoenvironmental research can play in addressing the
impacts of climate change. The broad objective of the project,
one of three regionally focused global change research initiatives co-ordinated by the Geological Survey of Canada, was
to enhance understanding of past environmental changes in
order to prepare for the potential biophysical impacts of
future climate change (Lemmen et al., 1993). It included
three primary research foci: 1) records of past climate and
hydrological changes, 2) understanding relationships
between climate and landscape processes, and 3) analysis of
landscape sensitivity. This volume presents the results of
work related to the first two of these.
The project's success resulted from the collaborative
approach taken by all participants. Basic funding for graduate
student research was provided through a competitive project
grant program, and supplemented by other sources of university funding. Project workshops were held in 1991 (Calgary),
1992 (Regina), and 1996 (Saskatoon). In total, the project
involved 35 researchers (including graduate students) from
Canada and the United States, contributing to 8 M.Sc. theses,
3 Ph.D. theses, and more than 50 journal publications.
While project research was restricted to geological
records and processes, it must be emphasized that the long
human history of the southern Prairies is another extremely
record of Holocene environmental change. As this important
subject is not represented in this volume, interested readers
are directed to Vickers (1986), Walker (1992), and Beaudoin
(1999).

STUDY AREA
Climate
While Captain Palliser's descriptions allow the Triangle to be
delineated on a map (Fig. I), the region is implicitly defined
on the basis of climate rather than geography. The most distinguishing characteristics are a strong annual moisture deficit and variable climate at all time scales. According to the
United Nations Environmental Programme (1992) climatic

classification, the area is dry subhumid to subhumid. Average


annual potential evapotranspiration exceeds average total
precipitation by 20 to more than 40%, with the driest portion
lying within southeastern Alberta and southwestern
Saskatchewan (Fig. 1). Recurrent drought is capable of producing true semiarid conditions over large areas (Wolfe,
1997). Drought has occurred in at least some portion of the
southern Prairies in virtually every decade of the twentieth
century (Chakravarti, 1976), with major regional droughts
occurring in the 1920s, the 1930s, 1961, and the 1980s. Consecutive drought years, such as 1936-1937 and 1987-1988,
brought about the most severe economic and environmental
impacts (Wheaton and Arthur, 1989).
The extreme climatic variability of the region is mainly a
function of its continental location in the lee of the Rocky
Mountains, and the interaction between Pacific and Arctic air
masses (Lemmen et al., 1998). The annual temperature range
of more than 80C is the largest in Canada (Hare and Thomas,
1979). The Canadian record-high air temperature of 45C
was recorded at Midale and Yellowgrass in southern
Saskatchewan on July 5, 1937. Winter record lows approach
-50C across most of the region (Wheaton, 1998). June is the
wettest month, typically accounting for about 25% of total
annual precipitation. Annual snowfall is relatively low
(70-180 cm), accounting for only about 30% of total annual
precipitation (Environment Canada, 1993).
The longest instrumental climate records from the Palliser
Triangle extend back slightly more than 100 years. Analysis
of long-term temperature trends reveals a statistically significant warming of 0.9"C during the period 1895-1991, within
which three distinct phases can be defined: a warming trend
from the late 1890s to the 1940s, a cooling trend from the
1940s to the 1970s, and a warming trend from the 1970s
through the 1990s (Fig. 2A; Gullett and Skinner, 1992). Precipitation does not show any statistically significant trend
over the century of instrumental record (Fig. 2B; Environment Canada, 1995), although Wolfe et al. (1998) noted a
cumulative drying trend since 1948, consistent with other
data suggesting increasing aridity (e.g. Judiesch and
Cutforth, 1996).

Biophysical environment
The Palliser Triangle is the northernmost extension of the
North American Great Plains, the vast grassland of the continental interior. It is a landscape of considerable topographic,
geological, and ecological diversity, all of which influence
land use (Fig. 3). The summary presented here is based on the
more detailed description of the geology, physiography,
soils, and vegetation of the central Palliser Triangle by
Lernmen et al. (1998).
Total relief in the Palliser Triangle exceeds 950 m. Rising
from 300 to 600 m above the adjacent plains, the Cypress
Hills (Fig. 1) include the highest point in southern Canada
between the Rocky and Torngat mountains (1465 m; see
Sauchyn, 1999). A number of uplands, including the Hand
Hills, Wood Mountain, Moose Mountain, and Turtle Mountain, dot the landscape (Fig. 1). Like most large-scale physiographic features on the southern Prairies, these uplands are

GSC Bulletin 534

Great Sand Hills of southwestern Saskatchewan (Fig. l), is


the largest contiguous dune area in southern Canada (David,
1977).

-Best-fit linear trend

-4.01

1900

$'
5P

100

.E

g-100
,-

n. -200

I
I

1920

1940

1960

" I '

1920

1940

1960

2000

'

-Five-year running mean


1900

1980

1980

2000

Year

Figure 2. Instrumental climate recordsfor the prairie region,


1895-1991, expressed as departures fronz 1951-1 980 mean:
A) tenzperature (adapted from Gullet and Skinner, 1992),
with solid line indicating best-fit linear trend (significantly
different from 0C a t the 95% confidence level) and
dashed lines indicating apparent shorter term trends;
B) precipitation (Environment Canada, 1995), with solid line
indicating five-year running mean.

largely bedrock controlled and most represent remnant surfaces of Late Tertiary fluvial erosion (Alden, 1932; Klassen,
1989). The Missouri Coteau, another major physiographic
feature of the Palliser Triangle, crosses the region as a northwest-trending escarpment rising from 50 to more than 250 m
above the plains to the east (Fig. 1).
Superimposed on bedrock-controlled physiographic elements are smaller scale landforms related to Pleistocene glaciation. With the exception of the highest parts of the Cypress
Hills and Wood Mountain upland, all of the region has been
glaciated (Klassen, 1989). Till is the dominant surficial
material, and there are extensive belts of ice-thrust features
and hummocky moraine, particularly along The Missouri
Coteau, the margins of the Cypress Hills, and the summits of
other uplands. Glaciolacustrine sediments, remnants of
extensive proglacial lakes, form the flat to rolling plains that
underlie the most productive agricultural soils of the region.
Glaciofluvial deposits, although limited in extent, are
extremely important surficial and shallow groundwater aquifers. Reworking of coarse glaciolacustrine deposits and
glaciofluvial sands by wind has produced significant areas of
sand dunes (see Muhs and Wolfe, 1999). One of these, the

Modern drainages strongly reflect the influence of glaciation. Glacial diversion of pre-existing drainages, plus large
tracts of hummocky terrain and a relatively dry climate, combine to produce large, internally drained basins covering
more than 25% of the region (Fisheries and Environment
Canada, 1978). Most through-flowing drainages have developed from a network of deglacial meltwater channels, many
of which are steep walled and incised up to 100 m. These features often form the most significant relief over large areas
(Kehew and Teller, 1994). The Palliser Triangle includes part
of the modern continental drainage divide (Fig. I), with runoff from parts of southernmost Alberta and Saskatchewan
flowing to the Gulf of Mexico via the Missouri, while the
remainder drains northeast to Hudson Bay.
The native vegetation cover of the Palliser Triangle is
referred to as 'northern mixed-grass prairie' (Risser et al.,
1981), or simply 'mixed prairie' (Coupland, 1950, 1961).
Considerable compositional variability results primarily
from the effects of relief and aspect on moisture availability.
Coupland (1950, 1961) utilized variation in seven major species to define all mixed prairie communities. Drought adaptations include apredominance of perennial, cool season forms
that begin growth in early spring and are dormant by
July (Risser et al., 1981). Since the onset of extensive
Eurocanadian settlement in the late nineteenth century, most
of this native grass cover has either been supplanted by cereal
crops or modified by grazing and the introduction of
non-native species. Of the ecological changes associated with
Eurocanadian settlement, Coupland (1961) deemed the elimination of fire as the single most significant event.
Uplands in the Palliser Triangle support forest cover, particularly along north-facing slopes and seepage areas. The
Cypress Hills support white spruce (Picea glauca), lodgepole
pine (Pinus contorta), trembling aspen (Populus
tremuloides), and balsam poplar (P, balsamifera), with
orographic precipitation providing critical moisture. The
Moose Mountain upland is distinguished by an extensive
paper birch (Betula papyrifera) population. Woody vegetation is also found in coulees, interdune areas, and deeply
incised river valleys where groves of cottonwood (Populus
acuminata, P. angustifolin, and P. deltoides) colonize alluvial flats.
The soils of the Palliser Triangle strongly reflect the subhumid climate and grassland vegetation of the region. Soils of
the Chernozemic Order dominate (Canada Soil Inventory,
1987a, b), and are by far the most important in terms of agricultural productivity. The region encompasses all of the
Brown Chernozemic Soil Zone (Fig. I), which was used by
Lemmen et al. (1998) as a working definition of the Palliser
Triangle. Dark Brown Chernozemic soils occur on the higher
elevations of plateaus and along the margin of the Brown Soil
Zone, while Chernozemic Black soils are found on the highest portions of the Cypress Hills. The differing colour of the
A-horizon of these soils reflects differences in organic matter
content (Rostad et al., 1993), which is in turn a reflection of

D.S. Lemmen and R.E. Vance

Figure 3. Geomorphic and land-use diversity in the Palliser Triangle: A) west block of the Cypress
Hills, Alberta, rising to 600 rn above the adjacent plains and supporting coniferous forest on
north-facing slopes; photograph courtesy of the Prairie Farm Rehabilitation Administration;
B) partially active sand d~irzesin the Great Sand Hills, Saskatchewan, the largest contiguous dune
occurrence in southern Canada; photograph by S.A. Wove; GSC 1999-042A; C) badlands in
Dinosaur Provincal Park, Alberta; small areas of badlands contribute large volumes of suspended
sediment to rivers; photograph courtesy of1.A. Campbell;D) and E) agriculture is the lifeblood of
the southern Prairies; lake plains gerzerally contain the most productive soils; grazing dominates
on sandy soils and in the driest portions of the Palliser Triangle, with dugouts providing critical
water for cattle; photographs courtesy of the Prairie Farm Rehabilitation Adnzinistration;
F ) recreation in Grasslands National Park, Saskatchewan; photograph by D.J. Sauchyn;
GSC 1999-042B.

GSC Bulletin 534

moisture availability. Significant areas of Solonetzic and


Regosolic soils also occur in the Palliser Triangle. Solonetzic
soils reflect a local or regional concentration of sodium in the
soil profile, and are commonly associated with areas of saline
soils. Regosols lack significant soil development, and are
regionally associated with three distinct environments: unstable sandy soils (including sand dunes), valley slopes of major
river systems, and areas of exposed bedrock.

RESULTS
Past climate and hydrological changes
An understanding of regional climatic variability, as well as
the timing and nature of past climatic changes, is essential to
any assessment of future climate-change impacts, and was
therefore the first major objective of the Palliser Triangle project. Although grasslands are potentially one of the more susceptible environments to climate change, they are also one of
the most difficult settings in which to obtain paleoenvironmental data (Barnosky et al., 1989). As a result, prior to the
initiation of this project, there was at best only a very rudimentary understanding of past climate dynamics in the
Palliser Triangle (Vance et al., 1992).
Tree-ling records are generally the most reliable, widely
distributed source of annual-resolution proxy climate data.
However, within the Palliser Triangle there is only one such
record available, a 3 12-yeardendroclimatic reconstruction of
precipitation from the Cypress Hills (Sauchyn and Beaudoin,

1998). This record indicates that some droughts of the eighteenth and nineteenth centuries were more severe than those
recorded in the historic (instrumental) climate record, and
that the most severe drought event occurred in the late 1700s.
Case and MacDonald (1995) arrived at similar conclusions
based on a 500-year dendroclimatic reconstruction of precipitation in the foothills of southern Alberta.
Despite the high resolution of the tree-ring data, the brief
duration of these records and the limited potential for additional sites are serious shortcomings in developing an understanding of regional paleoclimatic change. In an effort to
bridge this gap, Palliser Triangle project research focused on
millennium-scale reconstructions derived from
paleolimnological data (Vance, 1997). Such data provide not
only a proxy of past climate, but also a direct record of past
changes in surface-water quantity (water levels) and quality
(water chemistry). These paleohydrological reconstructions
are particularly important in the Palliser Triangle, where
water is considered to be the single most critical resource
issue with respect to impacts of future climate change
(Herrington et al., 1997).
Lakes are widespread throughout the Palliser Triangle,
and exhibit tremendous diversity in terms of hydrology, morphology, chemistry, and sedimentary processes (see detailed
review by Last (1999)). As well, many appear to respond predictably to environmental change, as shown by aerial photographs documenting recent water-level changes (Fig. 4).
Unfortunately, most lakes in the region are saline and ephemeral, and therefore contain not only a limited range of

Figure 4. Vertical aerial photographs showing dramatic drop in water level in


Antelope Lake, Saskatchewan (site 4, Fig. 5), between A) 1961 (NAPL A17302-26)
and B) 1991 (NAPL A27728-73).

D.S. Lemmen and R.E. Vance

biological proxy climate indicators but also discontinuous


sedimentary records. Hence, most Palliser Triangle lakes are
less than ideal candidates for paleoenvironmental studies.
However, lakes fed by shallow groundwater with significant
aquifer capacity are comparatively immune to frequent desiccation. and lakes situated on local uolands also tend to contain
continuous sedimentary records. Considerable effort was
therefore directed toward identification of appropriate
groundwater discharge and groundwater recharge sample
sites that formed transects across environmental gradients
(Vance and Last, 1994). General methods of core collection,
sampling, and analysis were described by Vance (1997).
~ d d i t i o n a ldetails can be found in publications of
site-specific studies (e.g. Vanceet al., 1997;Last et al., 1998).
Nine lakes were chosen for detailed paleolirnnological
study using multiple proxy indicators (Fig. 5 , Table 1;Vance,
1997). In addition, North Ingebrigt lake (Fig. 5) was the subject of detailed sedimentological and mineralogical studies
(Shang and Last, 1999), and plant macrofossil and
~edimentolo~ical
analyses of a prairie pothole on The Missouri Coteau (Andrews site, Fig. 5 ) provided a unique record
of late Pleistocene-early Holocene environments (Yansa,
1998; Yansa and Basinger, 1999; Aitken et al., 1999). Of the
nine principal study sites (Table l), four lie either on or along
the flanks of regional uplands (Elkwater, Harris, Kenosee,
and Max), two on The Missouri Coteau (Clearwater and Oro),
and three on the plains (Chappice, Antelope, and Killarney).
They reflect a range (Table 1) of water depth (<I to >9 m) and
salinity (freshwater (0.17 g - ~ - lto) hypersaline (165 g . ~ - l ) ) .
Although several lakes contain laminated sedimentary
sequences (Last and Vance, 1997), varves are not demonstrably present at any site. Geochronology was dependent
upon AMS radiocarbon dating of shoreline and terrestrial
plant macrofossils and 2 1 0 ~dating
b
(Vance, 1997).Sedimentary records of at least three of these study sites contain significant unconformities (Table 1).
A primary goal in the study of these sites was to integrate
multiple, independent proxies of environmental change in
order to develop a comprehensive evaluation of the role of
climate versus other limnological controls (Table 1). At

times, contradictions between various proxies are apparent,


as documented in a series of papers on Harris Lake (Sauchyn
and Sauchyn, 1991; Last and Sauchyn, 1993; Wilson et al.,
1997; Porter et al., 1999).Integrative approaches to resolving
these apparent contradictions were outlined in investigations
at Kenosee Lake (Vance et al., 1997) and Clearwater Lake
(Last et al., 1998; Leavitt et al., 1999). These approaches
resulted in reconstructions that include events and drivers
(principally hydrogeological and anthropogenic) that likely
could not be derived using any single indicator.
Detailed results from some sites remain unpublished, and
there has as yet been no opportunity to integrate individual
sites into a regional synthesis. Nonetheless, a preliminary
integration of results from these nine sites has allowed the
identification of three major hydroclimatic intervals within
the Holocene (Fig. 6; Vance et al., 1998).
Early Holocene (ca. 10 000-7000 BP)
Basal sediments in Harris, Clearwater, Oro, and Killarney
lakes all predate 9000 BP. These records indicate that the
early Holocene was an interval characterized by generally
high lake levels and freshwater conditions, although
short-term, low-water, more saline intervals are evident at all
sites. Rapidly changing conditions that do not appear synchronous between sites are interpreted to reflect a groundwater-dominated hydrology. Surficial and shallow aquifers,
fully charged following deglaciation, may have buffered
lakes from climate-driven changes in the early Holocene.
Since this groundwater effect appears to have been most pronounced on The Missouri Coteau (Clearwater Lake (Last et
al., 1998) and Oro Lake (Vance and Last, 1996)), an area of
extensive hummocky terrain, it may relate to the melting of
stagnant glacier ice thousands of years after retreat of the
active ice margin.
This strong groundwater influence makes definitive conclusions concerning early Holocene climate difficult. However, sedimentological and mineralogical evidence of arid
intervals do exist, such as 1) a saline to hypersaline phase

Figure 5. Principal study sites of the Palliser Triungle Global Change Project. Brown Chernozemic Soil Zone
is shaded. Open circles denote paleolimnological study sites, with numbers corresponding to those in Table 1.
Additional study sites are NI (North Ingebrigt lake) and AS (Anclrews site). Solid circles denote geomorphic
study sites.

GSC Bulletin 534

Table 1. Primary paleolimnological study sites of the Palliser Triangle Global Change Project. Site

numbers correspond to Figure 5. Analyses were conducted on multiple cores from each site.

Site

Latitude,
longitude

(references)

Maximum
deptha
(m)

Surface
TDS~
(g.~')

Core
lengthC
( 4

Basal
dated

730

0.32

121-165

8.7

7325 70

Yes

S, M, G, I, MA, P

1209

8.3

0.25-0.30

5.5

4940k 70

No

S, M, G, MA, 0

No
No

S, M, G, I, MA
S, M. G, MA, D,O, P

1850 70
(at 2.9 m)

No?

S, M, G, I, MA, A

3.2
3430 f 80
/ . i ' g 9 8 o * 7 o

NO
yes

Elevation
(m a.s.1)

Chappice

Unconformities

~nalyses~

Birks and Remenda.


1999)
Elkwater
(Vance and Last, 1994;
Vance. 19971

4g039'N,
11018'W

~arrls'
(Sauchyn and Sauchyn,
1991; Last and Sauchyn,
1993; Wilson et al., 1997;
Porter et ai., 1999; Wilson
and Smol, 1999)

Antelope
(Last and Vance, 1996;
Vance, 1997)
Clearwater
(Last et al., 1998; Leavitt
et al., 1999; Wilson and
Smol. 1999)
Or0
(Vance and Last. 1996;
Vance. 19971
Kenosee
IVance et al.. 19971

1
9

%ce and Last. 1994:


~ a n c e 1997)
,
Killarney
(Richmond and
Goldsborouah. 1999)

680

1
1
1

4g047'N,
10520'W
4 4 9 N
10218'W
49;'5N
',
100 09 W
49"l I'N,
99"42W

1
1

700

9.4

1( 11

0.87-1.21

6.4

17.5-40.2

4.3

2.07-3.63

0.26-0.34

735
665

1
I

No?

409011101

No

S,M.G.I.MA,O,A

3.8

No

S, M. G, MA. D 0 , A

Yes

S, M, G, I, MA, D, A

3330 k 70

9180 f 80

4.2

9420f230

S, M. G, I, MA, A
s, M, G,1, MA, D

8.1

0.44-0.64

5.9

490

3.5

1
/

1 1

S,M,G,I,MA,D,O,
PM

1
I
I

a Maximum depth recorded during water chemistry sampling.

Range of total dissolved solids observed in surveys conducted in August 1994, January 1995, and May 1995.
Refers to core collected from the central basin of each lake. A nearshore core from Clearwater Lake is included because it is much longer and older
than the core collected from the central basin.
In radiocarbon years, determined by accelerator mass spectrometry (AMS) dating of shoreline and terrestrial plant macrofossils. In most cases, dated
sample lies above base of core.
Analyses: S, sedimentology (includes particle size, water content, and organic matter); M, mineralogy; G , geochemistry; I, stable isotopes; MA,
macrofossils; D, diatoms; 0 ,ostracodes; A, algal pigments; P, pollen; PM, paleomagnetism.
Includes analyses of 9.6 m long core collected prior to start of the Palliser Triangle project.

(salinities approx. 20-70 g . ~ - lbetween


)
9800 and 8600 BP at
Clearwater Lake (Last et al., 1998); 2) an abrupt change from
freshwater conditions to a hypersaline, meromictic lake at
Oro Lake between ca. 9400 and 9000 BP (Vance and Last,
1996); and 3) a rapid transition from fresh to highly saline
conditions at ca. 10 000 BP at the Andrews site (Aitken et al.,
1999). It is possible that, at times, early Holocene climate
(associated with the insolation maximum) was as arid as, or
more arid than, during any subsequent interval. However,
since the hydrological setting was predominantly driven by
groundwater, high freshwater lake conditions dominated.

Mid-Holocene (ca. 7000-5000 BP)


The mid-Holocene was characterized by widespread low lake
levels in the Palliser Triangle. Only three of the study sites
(Chappice, Harris, and Oro lakes) contain sedimentary

records within the 7000-5000 BP interval. In each case, the


lake was significantly more saline than at present. Chappice
(Vance et al., 1992,1993) and Killarney (R.E. Vance, unpub.
data, 1997) lakes dried completely during this interval, the
only time this happened at either site during the Holocene.
The absence of sediments dating to this interval in other basins indicates that they were also completely dry at this time
(with the exception of Elkwater Lake, which was likely created by a landslide ca. 5000 BP). Present lake depths of 6 m
(Killarney Lake) to 9.5 m (Clearwater Lake) provide a rough
estimate of the magnitude of this mid-Holocene groundwater
decline (Lernmen et a1.,1997).Even at the end of this interval,
water levels in Kenosee Lake lay 5 m below present (Vance
et al., 1997). Remenda and Birlts (1999) suggested that the
actual mid-Holocene decline in regional water tables may
have been on the order of 6-15 m, based on depth of oxidized
sediments noted in hydrogeological studies.

D.S. Lemmen and R.E. Vance

Early Holocene (ca. 10 000 7000 BP)

Mid-Holocene (ca. 7000 5000 BP)

Late Holocene (ca. 5000 BP to present)

Water level

High Low

Salinity

2 Dry

Figure 6. Regional paleohydrological summary based on nine principal paleolimnological study


sites. Site numbers corresponcl with those in Table 1. Dashed line denotes eastern margin of Palliser
Triangle; areas of internal drainage are shaded. See text for discussion.

GSC Bulletin 534

The mid-Holocene was evidently an interval of severe


aridity across the southern Prairies. Based on the thickness of
distinctive laminated sequences indicative of drought events
in Chappice Lake, Vance et al. (1992) suggested that this
interval was one in which drought conditions (as defined by
historic climate) predominated for decades to perhaps centuries. Given the focus of the project -multiple proxy records
- quantitative climatic transfer functions were not developed as part of this study. However, previous estimates
derived from pollen records outside the Palliser Triangle in
Alberta and Manitoba suggest summer temperatures at this
time were about 2C warmer than present, and growing-season precipitation was 15% less (Vance et al., 1995).
These values are comparable to GCM projections for prairie
climate under an enhanced greenhouse gas (2 x C02) scenario
(Wheaton et al., 1992). Although the drivers of climatic
warming in the mid-Holocene were different (insolation)
than those used in future projections (greenhouse gases),
these paleoenvironmental results highlight the vulnerability
of prairie water resources to a warmer climate.
Late Holocene (ca. 5000 BP to present)
All study sites preserve a sedimentary record of this most
recent period of climatic and hydrological change. Rising
groundwater tables resulted in regional infilling of lake basins beginning about 5000 BP, roughly coincident with the
onset of Neoglaciation. Maximum water levels and freshwater conditions prevailed between about 3000 and 2000 BP. At
many lakes, evidence of fluctuations in water level and water
chemistry may correlate with well lcnown climatic events,
such as the Medieval Warm Period (ca. AD 900-1200) and
the Little Ice Age (ca. AD 1450-1850). Although limitations
in chronological control render such correlations tentative,
strong similarities in the records from several lakes suggest
regional climatic controls (Richmond and Goldsborough,
1999). In contrast, lakes in the Cypress Hills (Harris and
Elkwater lakes) remained fresh and relatively stable throughout this interval (Wilson and Smol, 1999). Another important
aspect of the late Holocene record is the impact of
Eurocanadian settlement and associated agricultural practices (Leavitt et al., 1999; Richmond and Goldsborough,
1999).In some basins, these anthropogenic impacts appear to
have overwhelmed any signal of recent climate fluctuations.
In general, the climate of the last 5000 years in the Palliser
Triangle was significantly more humid than during the preceding half of the Holocene. The relative importance of
increased precipitation versus decreased temperatures in
reducing regional aridity in the late Holocene remains
unknown. Besides the high water levels and freshwater conditions pervasive across the region, one of the strongest indications of cooler conditions was the proliferation of birch
(Betula papyrifera) on Moose Mountain upland during the
Little Ice Age (Vance et al., 1997). As important as the humid
extremes, however, is the magnitude of climatic variability
during the late Holocene. The thickness of laminated sediments at Chappice Lake suggests that droughts of the Medieval Warm Period lasted longer than those of the historic
(instmmental) climate record. This is particularly significant
since the major global controls on climate at this time were

essentially the same as at present, implying that such drought


events could occur in future, even in the absence of
anthropogenically induced climate change.
Synopsis
Among the common findings of the paleolimnological studies conducted as part of the Palliser Triangle project is that the
observed changes in fossil, mineralogical, and sedimentary
records reflect an intregration of climatic, hydrological, and
anthropogenic factors. Another commonality is the restriction imposed by limited chronological control (Vance, 1997).
Together, these factors make establishment of highresolution, long-term paleoclimatic records sensu strict0 difficult, and perhaps impossible, for this region. Nonetheless,
through the integration of multiple proxy records, the data
provide an important record of the paleohydrological
changes.
Other limitations of the paleolimnological record are also
emerging. For example, the drought of the late eighteenth
century, identified in the tree-ring record as the most severe of
the last 500 years, is not evident in the paleolimnological
record. This may relate to limited chronological control, but
also to the importance of antecedent conditions in controlling
the sensitivity of lake basins. The impact of drought will be
far less evident if it occurs during a generally humid interval,
owing to the buffer provided by fully charged shallow aquifers and the 'reservoir effect' of the lakes themselves. Lakes
at a low level with depleted local aquifers are much more
responsive to climatic change.
Beyond the basic framework of Holocene hydroclimatic
changes, a few fundamental conclusions must be emphasized
with respect to potential impacts of future climate change.
First, the historic (instrumental) climate record does not capture the full range of climate variability that can be expected
with present climate-forcing factors. This is particularly true
for aiTidity,since prior drought intervals appear to have been
far more severe than those of the twentieth centuly. Similar
conclusions have been reached by researchers studying the
American Great Plains (Laird et al., 1996; Woodhouse and
Overpeck, 1998). Second, because many lakes on the southern Prairies have experienced major changes in water levels
and water quality in response to past global changes, it is reasonable to conclude that future climate changes (resulting
from both anthropogenic forcing and climatic variability)
will also affect the region. But, perhaps most significantly, in
documenting the response of complex systems to past
changes in multiple controlling factors, these studies provide
the foundation for evaluating the potential impacts of future
climate change on surface and shallow groundwater
resources.

Landscape response to climate


The second major objective of the Palliser Triangle project
was to improve understanding of the relationships between
climate and geomorphic processes, in order to assess potential landscape response to future climate change. This work
builds upon the review of regional eolian, fluvial,

D.S. Lemmen and R.E.

mass-wasting, and soil redistribution systems presented by


Lemmen et al. (1998). Although climate is an important control in all of these geomorphic systems, they concluded that
eolian landscapes are the most sensitive to climate change.
Soil redistribution, however, has the greatest implication for
regional economic sustainability. While land-use decisions
and land-management techniaues are the most critical factors
influencing erosion of agricultural soils, there is high potential for significant soil loss in all cultivated areas under a climate that is warmer and drier than at present (Lernmen et al.,
1998).
u

Research on landscape response involved the following


approaches: geomorphic, stratigraphic, and geochronological analysis to determine past landscape changes; and monitoring of present rates of geomorphic processes that can be
related to instrumental climate data (Lemmen et al., 1993).
While both approaches have limitations when attempting to
assess future landscape change over time scales ranging from
decades to centuries (Sauchyn, 1999), they provide a means
of evaluating the climatic sensitivity of geomorphic systems
and a basis for estimating thresholds and response times
likely to be associated with future climate changes. Study
sites were clustered in the central, driest part of the Palliser
Triangle (Brown Chernozernic Soil Zone, Fig. 5 ) , assuming
that this is the most climatically sensitive area and therefore
most likely to show significant geomorphic response to climate change (Bull, 1991).
to assess the
In
changes on the prairie landscape, it must be noted that Late
Pleistocene glaciation was by far the single most important
event influencing recent regional landscape evolution
(Lernmen, 1998; Vreelten, 1999). Glaciation disrupted previously integrated drainages, significantly modified local topography through deposition of moraines and incision of
meltwater channels, and altered regional base levels through
glacio-isostatic adjustments (Klassen, 1989). Much of the
Palliser Triangle features a landscape still responding to
disequilibriums established more than 12 000 years ago.

Time

Late
Pleistocene

Early
Holocene

Vance

Hence landscape adjustments observed in the recent stratigraphic record largely reflect geomo~yhicresponse towards a
new dynamic equilibrium, upon which other factors, including climate change, are superimposed. The relative immaturity of the landscape is an important distinction between the
northernmost Great Plains and nonglaciated regions to the
south.
Variability in geomorphic activity throughout the
Holocene is presented schematically in Figure 7. This diagram represents a subjective assessment of activity relative to
the range observed for the Holocene as a whole, and is placed
in the context of the major hydroclimatic intervals discussed
previously. A more objective assessment of past geomorphic
activity, based on distribution of relevant radiocarbon dates,
has been developed for southeastern Alberta by Campbell
and Campbell (1997). The value of this approach is acknowledged, although it is not adopted here for the following reasons: l ) the limited number of absolute dates clearly related to
specific geomorphic events in the Palliser Triangle; 2) the
strong spatial clustering of what chronological data are available, with risk of local and event bias not representative of the
region; and 3) the bias of many stratigraphic records to either
recent or high-magnitude events (Sauchyn, 1999). In addition, significant response and relaxation times suggest that
simple chronological correlation between climatic and
geomorphic events does not demonstrate a causal relationship (Vreeken, 1999). Therefore, it must be emphasized that
~igur7
e presents a conceptual framework only, to be tested
and modified when a more extensive chronological database
is available,
In addition to documenting past geomorphic activity,
paleoenvironmental research makes it possible to assess the
relative importance of climate versus conditions internal to
geomorphic systems, as controls on regional geomorphic
activity (Table 2). Geomorphic systems that are near a
dynamicequilibrium willshow themostpredictableresponse
to climate change. In such cases, climate is herein deemed the
dominant factor controlling system response. However,
Late
Holocene

MidHolocene

Climate
Humid

Dealacial

Arid

System

Activitv

Figure 7.

Extreme

Fan formatlon

Schematic summary of relative geomorphic


activity thro~ighthe Holocene. Slzading denotes
intervals when activity was most regionally
pervasive. See textfor discussion. Note also compilation of Campbell and Campbell (1997),
based ciporz distribution of radiocarbon dates in
southern Alberta.

Fluvial

Mass-wasting

Extreme

rn

I---,

Extreme

Eolian

,,,D,U'?,aC?='?Y,z,

-m~r--

Loess deposition

Negligible

Soil erosion

---

Extreme
High
Low
Negligible

D a t e d e v e n t s )
Inferred

16

--

I-

I(
a

I
12

8
Time (x

1
4

l o 3 BP)

I
0

D.S. Lemrnen and R.E. Vance

between climate and dune activity is not simple, with dune


activation occurring as a threshold response controlled by
antecedent moisture conditions (Wolfe et al., 1998). The fact
that there has been a trend towards net dune stabilization
throughout the twentieth century (David, 1993; Wolfe and
Lernrnen, 1999), despite an overall warming trend and several severe droughts during this period, highlights the complexity of this system. Detailed chronological control through
optical dating (Huntley et al., 1985; Huntley and Lian, 1999)
suggests a response time of one to three decades for dune activation following extended severe drought (Table 2; David
et al., 1999). Relaxation times associated with dune stabilization may be even longer. For example, Wolfe and Lemmen
(1999) suggested that some of the present dune activity in the
Palliser Triangle is relict, relating to the regional activity of
the nineteenth century.
The predominance of agricultural lands in the Palliser
Triangle necessitates inclusion of soil erosion in any examination of regional landscape processes. Discussion of the
processes and rates of soil redistribution has been presented
by Pennock et al. (1995) and Lemmen et al. (1998). Rates of
wind and water erosion on bare soil are estimated to be two to
five orders of magnitude greater than those occurring on fully
vegetated pasture or native grassland (Evans, 1980; Coote,
1984). Thus, from the perspective of Holocene landscape
evolution, soil erosion was comparatively insignificant prior
to breaking of the prairie grassland by Eurocanadian settlers
(Fig. 7). In contrast, soil erosion today represents by far the
most regionally important geomorphic process, in terms of
both volume of sediment redistributed and associated economic costs.

response and relaxation times in many systems suggest that


some climate impacts will not be immediately apparent, but
may have considerable long-term consequences.

ACKNOWLEDGMENTS
The constructive comments of Steve Wolfe significantly
improved this manuscript. The authors would also like to
thank Trevor Robertson, Susan Ball, Murray Hay,
Mark Birchard, Jo-yi Wei, and Sonia Utting for cartographic,
field, and laboratory assistance while in the employ of the
Geological Survey of Canada.

REFERENCES
Aitken, A.E., Last, W.M., and Burt, A.K.
1999: The lithostratigraphicrecord of late Pleistocene-Holocene environmental change at the Andrews site near Moose Jaw, southern
Saskatchewan; in Holocene Climate and Environmental Change in
the Palliser Triangle: A Geoscientific Context for Evaluating the
Impacts of Climate Change on the Southern Canadian Prairies, (ed.)
D.S. Lemmen and R.E. Vance; Geological Survey of Canada,
Bulletin 534.
Alden, W.C.
1932: Physiography and glacial geology of eastern Montana and adjacent
areas; United States Geological Survey, Professional Paper 174,
133 p.
Barling, M.
1995: Postglacial evolution of Matzhiwin Creek, Alberta; M.Sc. thesis,
University of Alberta, Edmonton, Alberta, 83 p.
Barnosky, C. W.
al
and climate in the northwestern Great Plains
1989: ~ o s t ~ l a c ivegetation
of Montana:. Ouaternarv
Research,. v. 31. D. 57-73.
.
Beaudoin, A.B.
1999: What they saw: the climatic and environmental context for
Eurocanadiansettlementin Alberta; Prairie Forum, v. 24, p. 1 4 0 .
Birks, S.J. and Remenda, V.H.
1999: Groundwater inputs to a closed-basin saline lake, Chappice Lake,
Alberta; irz Holocene Climate and Environmental Change in the
Palliser Triangle: A Geoscientific Context for Evaluating the
Impacts of Climate Change on the Southern Canadian Prairies, (ed.)
D.S. Lemmen and R.E. Vance; Geological Survey of Canada,
Bulletin 534.
Bull, W.B.
1991: Geomorphic responses to climatic change; Oxford University
Press, New York, 326 p.
Campbell, C.
1998: Postglacial evolution of a fine-grained alluvial fan in the northern
Great Plains, Canada; Palaeogeography, Palaeoclimatology,
Palaeoecology, v. 139, p. 233-249.
Campbell, C. and Campbell, I.A.
1997: Calibration, review, and geomorphic implications of postglacial
radiocarbon ages in southeastern Alberta, Canada; Quaternary
Research, v. 47, p. 3 7 4 4 .
Campbell, I.A. and Evans, D.J.A.
1990: Glaciotectonism and landsliding in Little Sandhill Creek, Alberta;
Geomorphology, v. 4, p. 19-36.
Campbell, I.A., Rains, R.B., Shaw, J., and Evans, D.J.A.
1993: Glacial and Holocene geomorphologyof the Alberta Prairies; Field
Guide, Third International GeomorphologyConference, Hamilton,
Ontario, August 16-22, 1993, 92 p.
Canada Soil Inventory
1987a: Soil Landscapes of Canada - Alberta; AgricultureCanada, Publication 5237/B, 9 p. (plus tables, map at 1: 1 000 000 scale, and poster).
1987b: Soil Landscapes of Canada - Saskatchewan; Agriculture Canada,
Publication 5243/B, 9 p. (plus tables, map at 1:1 000 000 scale, and
poster).
. A

While anthropogenic activity is clearly the predominant


factor controlling soil redistribution (Table 2), wind and
water erosion of agricultural soils are closely related to
extreme climate events. Drought is a necessary precursor to
widespread wind erosion (Jones, 1991). The droughts of the
1930s clearly demonstrated that inappropriate land-use practices can lead to widespread destabilization of soils. In contrast, the droughts of the 1980s were climatically comparable,
but improved soil management resulted in far less severe
environmental impacts, despite severe economic impacts
(Wheaton and Arthur, 1989).
Research presented herein highlights a number of key
points with respect to assessing the potential geomorphic
impacts of future climate change. Perhaps most fundamental
is that climate is but one of several factors controlling
geomorphic response (e.g. Vreeken, 1999). Systems that are
still responding to past major disturbances are unlikely to
show a predictable response to climate change. Projections of
future impacts must be based on a thorough understanding of
system dynamics, including the importance of antecedent
conditions and threshold response (e.g. David et al., 1999).
While paleoenvironmental research makes an important contribution to this understanding, monitoring studies are also
important in providing direct measurement of processes
inferred for past geomorphic response (e.g. Wolfe and
Lemmen, 1999; Sauchyn and Nelson, 1999). Significant

GSC Bulletin 5 3 4
Case, R.A. and MacDonald, G.M.
1995: A clendrocli~naticreconstruction of annual precipitation on the
western Canadian Prairies since A.D. 1505 from Pirz~lsflexilis
James; Quaternary Research, v. 44, p. 267-275.
Chakravarti, A.K.
1976: Precipitation deficiency patterns in the Canadian Prairies, 1921 to
1970; Prairie Forum, v. 1, p. 95-1 10.
Christiansen, E.A. and Sauer, E.K.
1988: Age of the Frenchman Valley and associated drift south of the
Cypress Hills, Saskatchewan, Canada; Canadian Journal of Earth
Sciences, v. 25, p. 1703-1708.
Cohen, S.J.
1991: Possible impacts of climatic warming scenarios on water resources
in the Saskatchewan River sub-basin, Canada; Climatic Change,
v. 19, p. 291-3 17.
Coote, D.R.
1984: The extent of soil erosion in western Canada; in Proceedings of Second Annual Western Provincial Conference on Rationalization of
Water and Soil Research and Management: Soil Erosion and Land
Degradation; Saskatchewan Institute of Pedology, Saskatoon,
Saskatchewan, p. 34-49.
Coupland, R.T.
1950: Ecology of mixed prairie incanada; Ecological Monographs, v. 20,
p. 272-315.
1961: A reconsideration of grassland classification in the northern Great
Plains of North America; Journal of Ecology, v. 49, p. 135-167.
David, P.P.
1971: The Brookdale road section and its significance in the chronological
studies of dune activities in the Brandon Sand Hills of Manitoba;
Geological Association of Canada, Special Paper 9, p. 293-299.
1977: Sand dune occurrences of Canada: a theme and resource inventory
study of eolian landforms of Canada; Department of Indian and
Northern Affairs, Canada, National Parks Branch, Report, Contract
No. 74-230, 183 p.
1993: Great Sand Hills of Saskatchewan: an overview; irz Quaternary and
Tertiary Landscapes of Southwestern Saskatchewan and Adjacent
Areas, (ed.) D.J. Sauchyn; Canadian Plains Research Center,
University of Regina, Regina, Saskatchewan, p. 59-81.
David, P.P., Wolfe, S.A., Huntley, D.J., and Lemmen, D.S.
1999: Activity cycle of parabolic dunes based on morphology and chronology of Seward sand hills, Saskatchewan; in Holocene Climate
and Environmental Change in the PalliserTriangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.
de Boer, D.H.
1992: Hierarchies and spatial scale in process geomorphology: a review;
Geomorphology, v. 4, p. 303-318.
Environment Canada
1993: Canadian climate normals, 1961-1990: Volume 2, Prairie
Provinces; Minister of Supply and Services, 266 p.
1995: The state of Canada's climate: monitoring variability and change;
Environment Canada, Atmospheric Environment Service, SOE
Report 95-1, 52 p.
Evans, R.
1980: Mechanics of water erosion and their spatial and temporal controls:
an empirical viewpoint; i n Soil Erosion, (ed.) M.J. l r k b y and
R.P.C. Morgan; John Wiley and Sons, New York, p. 109-128.
Fisheries and Environment Canada
1978: Hydrological atlas of Canada; Fisheries and the Environment,
Ottawa, Ontario, 34 maps.
Gajeweski, K., Vance, R., Sawada, M., Fung, I., Gignac, D.,
Halsey, L., John, J., Maisongrande, P., Mandell, P., Mudie, P.J.,
Richard, P.J.H., Sherrin, A.G., Soroko, J., and Vitt, D.H.
in press: The cllmate of North America and adjacent ocean waters ca. 6 ka;
Canadian Journal of Earth Sciences.
Gullet, D.W. and Skinner, W.R.
1992: The state of Canada's climate: temperature change in Canada
1895-199 1; Environment Canada, Atmospheric Environment
Service, SOE Report No. 92-2, 36 p.
Hare, F.K. and Thomas, M.K.
1979: Climate Canada; John Wiley and Sons Canada Ltd., Toronto,
Ontario, 230 p. (second edition).

Herrington, R., Johnson, B., and Hunter, F.


1997: Responding to global climate change in the Prairies; Volume UI of
the Canada Country Study: Climate Impacts and Adaptation;
Environment Canada, 75 p. (plus append~ces).
Huntley, D.J. and Lian, O.B.
1999: Determining when a sediment was last exposed to sunlight by optical dating; i n Holocene Climate and Environmental Change in the
Palliser Triangle: A Geoscientific Context for Evaluating the
Impacts of Climate Change on the Southern Canadian Prairies, (ed.)
D.S. Lemmen and R.E. Vance; Geological Survey of Canada,
Bulletin 534.
Huntley, D.J., Godfrey-Smith, D.I., and Thewalt, M.L.W.
1985: Optical dating of sediments; Nature (London), v. 3 13, p. 105-107.
Jones, K.H.
1991: Drought on the Prairies; i n Symposium on the Smpacts of Climatic
Change and Variability on the Great Plains, (ed.) G. Wall;
University of Waterloo, Department of Geography, Occasional
Paper No. 12, p. 125-131.
Judiesch, D. and Cutforth, H.
1996: Weather trends at Swift Current: what's been going on for the last
century; Semiarid Prairie Agricultural Research Centre, Swift
Current, Saskatchewan, Research Newsletter, no. 20, p. 1-2.
Karl, T.R., Heim, R.R., Jr., and Quayle, R.G.
1991: The greenhouse effect in central North America: if not now, when?
Science, v. 251, p. 1058-1061.
Kehew, A.E. and Teller, J.T.
1994: History of late glacial runoff along the southwest margin of the
Laurentide Ice Sheet; Quaternary Science Reviews, v. 13,
p. 859-877.
Klassen, R.W.
1989: Quaternary geology of the southern Canadian Interior Plains; i n
Chapter 2 of Quaternay Geology of Canada and Greenland, (ed.)
R.J. Fulton; Geological Survey of Canada, Geology of Canada,
no. 1, p. 138-173 (also Geological Society of America, The
Geology of North America, v. K-1).
Laird, K.R., Fritz, S.C., Grimm, E.C., and Mneller, P.C.
1996: Century-scale paleoclimatic reconstruction from Moon Lake, a
closed basin lake In the northern Great Plains; Limnology and
Oceanography, v. 41, p. 890-902.
Laprise, R., Caya, D., Gigukre, IM.,
Bergeron, G., C8t6, H.,
Blanchet, J-P., Boer, G.J., and McFarlane, N.A.
1997: Climate and climate change in western Canada as simulated by the
Canadian Regional Climate Model; Atmosphere-Ocean, v. 36,
p. 119-167.
Last, W.M.
1999: Geolimnology of the Great Plains of western Canada; i n Holocene
Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lem~nenand
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Last, W.M. and Sauchyn, D.J.
1993: Mineralogy and lithostratigraphy of Harris Lake, southwestern
Saskatchewan, Canada; Journal of Paleolimnology, v. 9, p. 23-39.
Last, W.M. and Vance, R.E.
1996: Stop 12: Antelope Lake; irz Landscapes of the Palliser Triangle,
(ed.) D.S. Lemmen; Guidebook for the Canadian Geomorphology
Research Group Field Trip, Canadian Association of Geographers
1996 Annual Meeting, Saskatoon, Saskatchewan, p. 4 2 4 3 .
1997: Bedding charactcristics of Holocene sediments from salt lakes of
the northern Great Plains, western Canada; Journal of
Paleolimnology, v. 17, p. 297-318.
Last, W.M., Vance, R.E., Wilson, S., and Smol, J.P.
1998: A multi-proxy record of early Holocene hydrologic change on the
northern Great Plains of southwestern Saskatchewan, Canada; The
Holocene, v. 8, p. 503-520.
Leavitt, P.R., Vinebrooke, R.D., Hall, R.I., Wilson, S.E., Smol, J.P.,
Vance, R.E., and Last, W.M.
1999: Multiproxy record of prairie lake response to climate change and
human activity, Clearwater Lake, Saskatchewan; irz Holocene
Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and
R.E. Vance; Geological Survey of Canada, Bulletin 534.

D.S. Lemmen and R.E. Vance


Lemmen, D.S.
1998: Geomorphic response to Holoceneclimaticchanges on the southern
Canadian Prairies; Geological Society of America, Annual
Meeting, Abstracts with Programs, v. 30, no. 7, p. A-251.
Lemmen, D.S., Dyke, L.D., and Edlund, S.A.
1993: The Geological Survey of Canada's Integrated Research and
Monitoring Area (IRMA) projects: a contribution to Canadian
global change research; Journal of Paleolimnology, v. 9, p. 77-83.
Lemmen, D.S., Vance, R.E., Campbell, I.A., David, P.P., Pennock,
D.J., Sauchyn, D.J., and Wolfe, S.A.
1998: Geomorphic systems of the Palliser Triangle, southern Canadian
Prairies: description and response to changing climate; Geological
Survey of Canada, Bulletin 521,72 p.
Lemmen, D.S., Vance, R.E., Wolfe, S.A., and Last, W.M.
1997: Impacts of future climate change on the southern Canadian Prau-ies:
a paleoenvironmental perspective; Geoscience Canada, v. 24,
p. 121-133.
Mann, M.E., Bradley, R.S., and Hughes, M.K.
1998: Global-scale temperature patterns and climate forcing over the past
six centuries; Nature, v. 392, p. 779-787.
Muhs, D.R. and Wolfe, S.A.
1999: Sand dunes of the northern Great Plains of Canada and the United
States; in Holocene Climate and Environmental Change in the
Palliser Triangle: A Geoscientific Context for Evaluating the
Impacts of Climate Change on the Southern Canadian Prairies, (ed.)
D.S. Lemmen and R.E. Vance; Geological Survey of Canada,
Bulletin 534.
Nemanishen, W.
1998: Droughtin thePailiserTriangle(aprovisional primer); Prairie Farm
Rehabilitation Administration, Agriculture and Agrifood Canada,
58 p.
Nkemdirim, L.C. and Purves, H.
1994: Estimating the potential impact of climate change on streamflow in
the Oldman River basin, Alberta: an analogue approach; Canadian
Water Resources Journal, v. 19, p. 141-155.
O'Hara, S.L. and Campbell, I.A.
1993: Holocene geomorphology and stratigraphyof (he lower Falcon valley, Dinosaur Provincial Park, Alberta, Canada; Canadian Journal
of Earth Sciences, v. 30, p. 1846-1852.
Pennock, D.J., Lemmen, D.S., and de Jong, E.
1995: Cesium-137-measured erosionrates for soils of five parent-material
groups in southwestern Saskatchewan; Canadian Journal of Soil
Science, v. 75, p. 205-210.
Porter, S.C., Sauchyn, D.J., and Delorme, L.D.
1999: The ostracode record from Harris Lake, southwestern
Saskatchewan: 9200 years of local environmental change; Journal
of Palcolimnology,v. 21, p. 35-44.
Rains, R.B., Burns, J.A., and Young, R.R.
1994: Postglacial alluvial terraces and an incorporated bison skeleton,
Ghostpine Creek, southern Alberta; Canadian Journal of Eal.th
Sciences, v. 31, p. 1501-1509.
Remenda, V.H. and Birks, S.J.
1999: Groundwaterin the Palliser Triangle: an overview of its vulnerability and potential to archive climate information; in Holocene
Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lem~nenand
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Richmond, K.-A. and Goldsborough, L.G.
1999: Late Holocene paleolimnology of Killarney Lake, Manitoba; iiz
Holocene Climate and Environmental Change in the Palliser
Triangle: A Geoscientific Context for Evaluating the Impacts of
Climate Change on the Southern Canadian Prairies, (ed.)
D.S. Lemmen and R.E. Vance; Geological Survey of Canada,
Bulletin 534.
Risser, P.G., Birney, E.C., Blocker, H.D., May, S.W., Parton, W.J.,
and Weins, J.A.
1981: The True Prairie Ecosystem; Hutchinson Ross Publishing Company, Stroudsburg, Pennsylvania, 557 p.
Rostad, H.P.W., Bock, M.D., Krug, P.M., and Stushnoff, C.T.
1993: Organic matter content of Saskatchewan soils; Saskatchewan
Institute of Pedology, Saskatoon, Saskatchewan, Publication
No. MI 14.

Sauchyn, D.J.
1999: Geomorphology of the western Cypress Hills: climate, process,
stratigraphy and theory; in Holocene Climate and Environmental
Change in the Palliser Triangle: A Geoscientific Context for Evaluating the Impacts of Climate Change on the Southern Canadian
Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological Survey of
Canada, Bulletin 534.
Sanchyn, D.J. and Beaudoin, A.B.
1998: Recent environmental change in the southern Canadian Plains; The
Canadian Geographer, v. 42, p. 337-353.
Sauchyn, D.J. and Lemmen, D.S.
1996: Impacts of landsliding in the western Cypress Hills, Saskatchewan
and Alberta; in Current Research 1996-B; Geological Survey of
Canada, p. 7-14.
Sauchyn, D.J. and Nelson, H.L.
1999: Origin and erosion of the Police Point landslide, Cypress Hills,
southeastern Alberta; in Holocene Climate and Environmental
Change in the Palliser Triangle: A Geoscientific Context for Evaluating the Impacts of Climate Change on the Southern Canadian
Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological Survey of
Canada, Bulletin 534.
Sauchyn, M.A. and Sauchyn, D.J.
1991: A continuous record of Holocene pollen from Harris Lake, southwestern Saskatchewan, Canada; Palaeogeography,
Palaeoclimatology,Palaeoecology, v. 88, p. 12-23,
Shang, Y. and Last, W.M.
1999: Mineralogy, lithosttatigraphy, and inferred geochelnical histoly of
North Ingebrigt lake, Saskatchewan; in Holocene Climate and
Environmental Change in the Palliser Triangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.
Smit, B., Burton, I., Klein, J.T., and Street, R
in press: The science of adaptation: a framework for assessment; Mitigation
and Adaptation Strategies for Global Change.
Spry, I.
1968: The papers of the Palliser Expedition 1857-1860; The Champlain
Society, Toronto, Ontario, 694 p.
Thomson, S. and Morgenstern, N.R.
1978: Landslides in argillaceous rock, Prairie Provinces, Canada; in
Rockslides and Avalanches 2, (ed.) B. Voight; Elsevier Scientific
Publishing, New York, Developments in Geotechnical Engineering, 14B, p. 515-540.
United Nations Environmental Programme (UNEP)
1992: World Atlas of Desertification;Edward Arnold, Sevenoaks,United
Kingdom, 69 p.
Vance, R.E.
1997: The Geological Survey of Canada's Palliser Triangle Global
Change Project: a multidisciplinary geolimnological approach to
predicting potential global change impacts on the northern Great
Plains; Journal of Paleolimnology, v. 17, p. 3-8.
Vance, R.E. and Last, W.M.
1994: Paleolimnology and global change on the southern Canadian Prairies; in Current Research 1994-B; Geological Survey of Canada,
p. 49-58.
1996: Stop 4: Oro Lake; in Landscapes of the Palliser Triangle, (ed.) D.S.
Lemmen; Guidebook for the Canadian Geomorphology Research
Group Field Trip, Canadian Association of Geographers 1996
Annual Meeting, Saskatoon, Saskatchewan, p. 26-27.
Vance, R.E., Beaudoin, A.B., and Luckman, B.H.
1995: The paleoecologicalrecord of 6 ka BP climate in the Canadian prairieprovinces; GCographie physique et Quaternaire,v. 49, p. 8 1-98.
Vance, R.E., Clague, J.J., and Mathewes, R.W.
1993: Holocene paleohydrology of a hypersaline lake in southeastern
Alberta; Journal of Paleolimnology, v. 8, p. 103-120.
Vance, R.E., Last, W.M., and Lemmen, D.S.
1998: The rain (or lack thereof) on the plain: putting the Canadian Great
Plains on the paleoclimate map; First International Geosphere
Biosphere Program PAGES (Past Global Changes) Open Science
Meeting, Abstracts Volume, p. 128.
Vance, R.E., Last, W.M., and Smith, A.J.
1997: Hydrologic and climatic implications of a multidisciplinary study
of late Holocene sediment from Kenosee Lake, southeastern
Saskatchewan,Canada; Journal of Paleolimnology,v. 18, p. 1-29.

GSC Bulletin 5 3 4
Vance, R.E., Mathewes, R.W., and Clague, J.J.
1992: 7000-year record of lake-level change on the northern Great Plains:
a high-resolution proxy ofpast climate; Geology, v. 20,p. 879-882.
Vickers, J. R.
1986: Alberta plains prehistory: a review; Archaeological Survey of
Alberta, Alberta Culture, Edmonton, Alberta, Occasional Paper
No. 27, 139 p.
Vreeken, W.J.
1996: A chronogram for postglacial soil-landscape change from the
Palliser Triangle, Canada; The Holocene, v. 6, p. 4 3 3 4 3 8 .
1999: Geomorphic surfaces and postglacial landscape evolution of the
Maple Creek basin, Saskatchewan; in Holocene Climate and
Environmental Change in the Palliser Triangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.
Walker, E.G.
1992: The Gowen sites: cultural responses to climatic warming on the
northern plains; Archaeological Survey of Canada, Canadian
Museum of Civilization, Ottawa, Mercury Series Paper 145,208 p.
Wheaton, E.E.
1998: But It's a Dry Cold! Weathering the Canadian Prairies; Fifth House
Ltd., Calgary, Alberta, 185 p.
Wheaton, E.E. and Arthur, L.M.
1989: Executive summary; irz Environmental and Economic Impacts of
the 1988 Drought: With Emphasis on Saskatchewan and Manitoba,
(ed.) E.E. Wheaton and L.M. Arthur; Saskatchewan Research
Council, Saskatoon, Publication No. E-2330-4-E-89, p. iii-xxiv.
Wheaton, E.E., Wittrock, V., and Williams, G.D.V.
1992: Saskatchewan in a warmer world: preparing for the future;
Saskatchewan Research Council, Saskatoon, Publication
No. E-2900-17-E-92, 13 p.
Wilson, S.E. and Smol, J.P.
1999: Diatom-based salinity reconstructions from Palliser Triangle lakes:
a summary of two Saskatchewan case studies; in Holocene Climate
and Environmental Change in the Palliser Triangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.

Wilson, S.E., Smol, J.P., and Sauchyn, D.J.


1997: A Holocene paleosalinity diatom record from southwestern
Saskatchewan, Canada: Harris Lake revisited; Journal of
Paleolimnology, v. 17, p. 23-31.
Wolfe, S.A.
1997: Impact of increased aridity on sand dune activity in the Canadian
Prairies; Journal of Arid Environments, v. 36, p. 4 2 1 4 3 2 .
Wolfe, S.A. and Lemmen, D.S.
1999: Monitoring of sand dune activity in the Great Sand Hills region,
southwestern Saskatchewan; in Holocene Climate and Environmental Change in the Palliser Triangle: A Geoscientific Context for
Evaluating the Impacts of Climate Change on the Southern
Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological
Survey of Canada, Bulletin 534.
Wolfe, S.A. and Nickling, W.G.
1997: Sensitivity of eolian processes to climate change in Canada;
Geological Survey of Canada, Bulletin 421, 30 p.
Wolfe, S.A., Huntley, D.J., and Ollerhead, J.
1995: Recent and late Holocene sand dune activity in southweste~n
Saskatchewan; in Current Research 1995-B; Geological Survey of
Canada, p. 131-140.
Wolfe, S.A., Huntley, D.J., Sauchyn, D.J., David, P.P., Ollerhead, J.,
and MacDonald, G.
1998: Geochronologic evidence for widespread dune activity induced by
late 18Ih century dryness, Great Sand Hills, southwestern
Saskatchewan; Geological Society of America, Annual Meeting,
Abstracts with Programs, v. 30, no. 7, p. A-217.
Woodhouse, C.A. and Overpeck, J.T.
1998: 2000 years of drought variability in the central United States;
Bulletin of the American Meteorological Society, v. 79,
p. 2693-2714.
Yansa, C.H.
1998: Holocene paleovegetation and paleohydrology of a prairie pothole
in southern Saskatchewan, Canada; Journal of Paleolimnology,
v. 19, p. 4 2 9 4 4 1 .
Yansa, C.H. and Basinger, J.F.
1999: A postglacial plant macrofossil record of vegetation and climate
change in southern Saskatchewan; in Holocene Climate and
Environmental Change in the Palliser Triangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.

Geolimnology of the Great Plains


of western Canada
William M. ~ a s t l
Last, W.M. 1999: Geolimnology of the Great Plains of western Canada; in Holocene Climate and
Environmental Change in the Palliser Triangle: A Geoscientific Context for Evaluating the
Impacts of Climate Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and
R.E. Vance; Geological Survey of Canada, Bulletin 534, p. 23-55.

Abstract: Despite a long history, lacustrine research on the western Canadian Great Plains has only
recently focused on understanding the complex, interactive physical, chemical, and biological processes
that characterize modern lakes. These lakes exhibit tremendous diversity in terms of hydrology, morphology, chemical composition, and sedimentary processes. Although many basins exhibit playa characteristics, the region also contains several of North America's largest lakes. Salinity and ionic composition show
great diversity, with virtually every type of water chemistry represented. Associated with this large range in
brine chemistry is an equally diverse assemblage of endogenic and authigenic minerals. Although investigation of the Holocene stratigraphic records in these basins is in its infancy, some approaches, such as deciphering paleochemistry and paleohydrology from the endogenic mineral record and isotopic composition,
have proven to be successful. Challenges for both fundamental and applied researchers are to integrate, on a
regional basis, the sedimentological and geochemical complexities exhibited by modern lakes with the preserved stratigraphic records.
RCsum6 : En dCpit d'une longue histoire, la recherche limnologique dans les grandes plaines
canadiennes de 1'Ouest ne s'attache que depuis peu ?I la comprChension des processus physiques, chimiques
et biologiques interactifs complexes qui caractCrisent les lacs modernes. Ces lacs tCmoignent d'une
remarquable diversit6 quant ?I leur hydrologie, leur morphologie, leur composition chirnique et leurs
processus skdimentaires. M&mesi de nombreux bassins prdsentent des caractgristiques de playas, cette
rCgion renferme plusieurs des plus grands lacs de 1'AmCrique du Nord. La salinitC et la composition ionique
presentent Cgalement une diversit6 marquk, pratiquement chaque type de chimie des eaux Ctant reprCsent6.
On trouve associt ?I ce large Cventail de chimie des saumures un assemblage tout aussi variC de min6raux
endogbnes et authigbnes. Bien que les recherches sur les donntes stratigraphiques holocbnes de ces bassins
soient encore embryonnaires, certaines approches, notamment le dCchiffrage de la palCochimie et de la
palCohydrologie ?I partir des donnCes sur les minCraux endogbnes et sur les compositions isotopiques, se
sont av6rkes fructueuses. L'intCgration sur une base regionale des complexitCs ~Cdimentologiqueset
gCochimiques que prCsentent les lacs contemporains et des donnCes stratigraphiques qui ont CtC conservCes
constitue un dCfi pour les chercheurs en sciences tant fondamentales qu'appliqu6es.

Department of Geological Sciences, University of Manitoba, Winnipeg, Manitoba R3T 2N2

GSC Bulletin 534

INTRODUCTION
...in the central part of the continent there is a region, desert,
or semi-desert in character, which can never be expected to
become occupied by settlers...
- Palliser, 1862

The stress that climate imposes on the southern Canadian


Prairies, clearly recognized by Captain John Palliser, has
been amplified in the present century as the region evolved
into one of the world's most important agricultural regions.
Evaluation of the future sustainability of the Palliser Triangle, including possible impacts associated with changing climate, depends upon a thorough knowledge of the sensitivity
and dynamics of both natural and anthropogenic systems. T o
enhance our understanding of the former, a primary objective
of the Geological Survey of Canada's Palliser Triangle
Global Change Project was to decipher the timing and severity of postglacial climatic and hydrological changes, and their
impact on the prairie landscape (Lemmen et al., 1993). Investigations of Holocene stratigraphic records preserved in the
lakes of the region are pivotal in accomplishing this objective
(Lernmen et al., 1997; Vance, 1997). These basins offer an
outstanding opportunity to study past environmental conditions, owing to their great diversity in morphology, chemistry, and depositional processes. Furthermore, the large
number of lacustrine basins and their presence over such a
broad area will ultimately allow integration of site-specific
studies in a variety of hydrological and geomorphic settings,
such as presented elsewhere in this volume, in order to gain a
regional perspective of environmental changes.
Geolimnology is the study and interpretation of physical,
geochemical, and hydrogeological processes in lakes and
sedimentological records of lacustrine basins. Although the

sedimentological and geochemical examination of lakes had


its origin during the latter part of the nineteenth century, progress throughout most of the twentieth century was slow.
However, beginning in the 1970s, the recognition that lake
sediments provide a source of valuable industrial minerals
and fossil fuels, as well as the increased use of the inorganic
components of lake sediments to monitor pollution, decipher
environmental changes, and deduce past climatic and hydrological conditions, dramatically increased interest in lacustrine geological processes and lake deposits. The past decade
has seen rapid advance in our understanding of the physical
and chemical processes operating in lakes and how these processes apply to the stratigraphic sequences preserved in lacustrine basins (e.g. Anad6n et al., 1991; GierlowskiKordesch and Kelts, 1994; Renaut and Last, 1994; Talbot and
Allen, 1996). Indeed, it is now generally accepted that probably no other continental setting has as much to offer in terms
of potentially significant contributions to the earth sciences as
the lake environment.
The objective of this paper is to introduce the wide assortment of modern lacustrine environments in the northern
Great Plains of western Canada (Fig. I), and to provide an
overview of recent advances in order to establish a framework for future geolimnological and paleolimnological
research efforts in this region. Assimilation of over 120 years
of research on these lakes is an onerous task, even for the most
dedicated reader. In this paper I have placed most emphasis
on saline lakes and saline lake sediments, on extant lakes and
their sediment records, and on the physical, mineralogical,
and geochemical components, processes, and stratigraphies
in these basins. After a brief summary of the physical and climatic setting of the region, I review the major morphological
types of lake basins and their origin. This is followed by a discussion of the chemical composition and properties of the

Figure 1. Location map of the northern Great Plains of western Canada. Cross-hatched shading
indicates areas of internal drainage. Inset map shows the location of the three Prairie Provinces in
Canada.

W.M. Last

waters, and a synopsis of the major theories regarding the origin of the salts in the lakes. A description of the modern sedimentary processes and resulting sedimentary facies is
followed by an outline of the major geolimnological tools
used by paleolimnologists to help understand the history of
these lakes. Finally, 1 offer some speculative comments on
the future of geolimnological research in the Great Plains of
western Canada. Aspects of economic use of these lakes are
presented in Appendix A, and Appendix B offers information
on brine salinity and measurement.
The work of scientists from fields other than geology,
geochemistry, and geography needs to be examined for a
complete understanding of the lakes in this region. Overviews of the biological aspects are provided in Rawson and
Moore (1944), Nol-thcote and Larkin (1963), Haynes and
Hammer (1978), Hammer (1981,1984,1986), Hammer et al.
(1983), Waite (1986), and Adams (1988). The sedimentology, chronology, history, and development of extinct
(mainly glacial and proglacial) lakes and wetlands also form a
very important part of the overall scientific picture of the
northern Great Plains. These are summarized in Christiansen
(1979), Klassen (1989, 1994), Beaudoin (1993), and Teller
and Kehew (1994). With the notable exception of Lake
Winnipeg (Brunskill and Graham, 1979; Todd et al., 1996,
1998), there has been little geolimnological research on the
few but interesting freshwater basins.

SETTING
A great untimbered, level, dried-up sea of land.. .
- Wilby, 1914

Although the physical, geological, and climatic diversity


of the northern Great Plains has been summarized elsewhere
(e.g. Hammer, 1978a; Wilson and Dijks, 1993; Mossop and
Shetsen, 1994; Teller et al., 1996; Lemmen et al., 1998), the
following synthesis emphasizes the most important characteristics of the region that impart distinctive, and often
unique, characteristics to geolimnological features.

Smith, 1993) and, by the early Quaternary, a mature dendritic


drainage pattern had been established over much of the northern Great Plains (Stalker, 1961; Christiansen, 1967b; Bird,
1980).In general, this ancestral pattern is reflected by today's
streams, except that much of the upper Missouri River flowed
northeastward into Hudson Bay rather than into the
Mississippi River basin (Meneley et al., 1957; Bluemle,
1991).
The bedrock is mantled by unconsolidated Quaternary
sediment which is over 300 m thick in places (Klassen, 1989).
These deposits consist of till, glaciofluvial sand and gravel,
and lacustrine silt and clay. During deglaciation, meltwater
from the retreating Laurentide Ice Sheet carved numerous
ice-marginal channels and spillways into this sediment and
bedrock (Christiansen, 1979; Kehew and Lord, 1989; Kehew
and Teller, 1994). Although now abandoned or buried under
more recent sediment, these valleys often form modern lake
basins. The hydrodynamic properties of Quaternary fill in the
valleys strongly influence the character and composition of
Holocene sediments in these lakes by controlling the direction of flow and quantity of groundwater discharge (Witkind,
1952; Rueffel, 1968a; Wood and Sanford, 1990; Donovan,
1992; Donovan and Rose, 1994).

Physiography
Within the northern Great Plains physiographic province
(Bird, 1980), locally separate and distinct subprovinces can
be identified. The Manitoba Lowland is a low (<300 m elevation) wetland area extending westward from the Precambrian
Shield to the Manitoba Escarpment. This area contains the
largest lakes in the Great Plains: Winnipeg, Manitoba,
Winnipegosis, Cedar, and Dauphin. All are remnants of glacial Lake Agassiz and have large drainage basins and generally high sedimentation rates, which result in a long and
exceedingly complex sedimentary record that reflects
glacio-isostasy, evolving landscapes, variable fluvial inputs,
and regional climate fluctuations.
Extending westward from about the Manitoba border is
the Saskatchewan Plains region, sometimes referred to as the

Geology
The Canadian Great Plains are underlain by nearly horizontal
Phanerozoic sedimentary rocks up to 5000 m thick. The
Paleozoic section consists mainly of a series of stacked carbonate-evaporite cycles, whereas the overlying Mesozoic
and Cenozoic bedrock is dominantly a sand-shale sequence.
Dissolution of the highly soluble Paleozoic evaporites has
modified the relatively simple shuctural relationships of the
flat-lying formations and has created collapse structures over
much of the area(Christiansen, 1967a, 1971). Several authors
have suggested that this evaporite dissolution has provided a
source of ions for the many salt lakes of the region (Cole,
1926; Sahinen, 1948; Grossman, 1949, 1968; see also
reviews in Last (1988) and Murphy (1996)).
The bedrock surface has been strongly modified by
preglacial erosion. Subaerial erosion during the Late Tertiary
removed up to 900 m of sediment in some areas (Leckie and

'second prairie level'. This is a large area of gentle relief,


between about 450 and 700 m a.s.l., containing a very large
number of mainly small and shallow lakes. Estimates range
from 4 to 10 million lakes and wetlands in this region (Gollop,
1963; Adams, 1988) and densities as high as 90-120
lake~.km-~
in some localities (Last, 1989a; Adams et al.,
1992). The Saskatchewan River Lowlands in the east give
way to the greater topographic diversity of the Central
Saskatchewan Plains and various uplands areas.
Separating the second prairie level from the western portion of the geomorphic province is The Missouri Coteau and
its eastward-facing edge, the Missouri escarpment. The Missouri Coteau is a distinct, 50-100 krn wide band of knob-andkettle topography that extends for over 1200 km through the
Great Plains from central South Dakota northwestward into
west-central Saskatchewan (Clayton and Freers, 1967). The
terrain in this region is rougher and the elevation
(550-1400 m) greater than in the Central Saskatchewan

GSC Bulletin 534


Plains to the east. 'Badlands' topography has developed in
some areas, with local relief being as much as 150 m. The
Coteau contains many saline to hypersaline lakes.
The Alberta Plains, or third prairie level, extend westward
from western Saskatchewan to the foothills of the Rocky
Mountains. Although this western part of the Great Plains
of the
contains
fewer lake basins,
best-studied Holocene lacustrine stratigraphic sequences
occur here.

season is moderate to high and mainly from the west and


southwest (see Wolfe and Lernmen, 1999). In addition to
greatly aiding evaporation, wind plays a major role in current
and wave generation and, hence, sediment deposition and
erosion (Kenny, 1985). Wind has also been shown to cause
significant local variation in sedimentary facies patterns
within basins, and can be an important agent for transporting
elastic sediment and salt into and out of lakes (Last, 1989a;
Donovan, 1994; Wood and Sanford, 1995).

Hydrology

Climate
The most distinguishing climatic characteristics of the
Canadian Great Plains are its extreme variability in temperature and its strong annual moisture deficit. Mean daily temperatures during January of about -18C and during July of
about 19C (Canadian National Committee for the International Hydrologic Decade, 1978) do not adequately convey
the temperature variability experienced on a diurnal, seasonal, and inter-annual basis. This variability has a significant
impact on many chemical and physical aspects and processes
of the lakes in the region. Temperature, as a primary control
of evaporation, is also a contributing factor to the high annual
moisture deficit. The region receives about 40 cm of precipitation per year (Fig. 2), but as much as 1.5 m of water can be
lost annually through evaporation from open water bodies
(Canadian National Committee for the International
Hydrologic Decade, 1978). This negative moisture balance
helps to impart and control the characteristically high salinities of water in most lakes of the region.
Wind is
in dictating the processes 'persting in lakes.
not
lakes form a winter
ice cover, the average wind speed during much of the ice-free

The single most significant factor influencing the nature, distribution, and sedimentary characteristics of the lakes is probably the presence of large areas of internal drainage in the
northern Great Plains (Fig. 1). These basins form one of the
largest and best-studied areas of endorheic drainage in North
America (Langham, 1970; Rutherford, 1970; Whiting, 1974,
1977,1986; Last, 1984a). The lack of integrated drainagepatterns over such large regions makes precise definition of
watersheds and drainage divides difficult (Lernmen et al.,
1998). Those areas not characterized by closed basins are
drained by three major river systems: the Saskatchewan, the
Qu'Appelle-Assiniboine, and the Milk. The first two of these
drainages flow to Hudson Bay, whereas the Milk River and its
tributaries drain the southernmost parts of Saskatchewan and
Alberta to the Gulf of Mexico via the Missouri River system.
Groundwater plays a pivotal role in the regional
geolimnology. As summarized elsewhere in this volume
(Remenda and Birks, 1999), subsurface water compositions
in the region are of several main types (see also overviews in
Rutherford (1967), Brown (1967), Freeze (1969), Tokarsky
(1985, 1986). Lennox et a l (1988). Pupp et al. (1991),
Betcher et al. (1995).

Figure 2. Mean annual precipitation (solid lines) and evaporation (dashed lines) in milliinetres for
the northern Great Plains (modified froin Canadian National Committee for the International
Hydrologic Decade, 1978, Fig. 25).

W.M. Last

GENESIS AND MORPHOLOGY


Lakes arise from phenomena that are almost entirely geologic in nature. Once formed, they are doomed. Because of
the concave nature of basins, there is a compulsory trend
toward obliteration as they fill with sediments...
- Cole, 1975, p. 9
Most lakes of the Canadian Great Plains are of glacial origin, as all but the highest reaches of the Cypress Hills and
Wood Mountain upland were ice covered during the Late
Wisconsinan (Dyke and Prest, 1987; Klassen, 1989). Comparatively few lake basins can be directly attributed to active
glacial ice, but many are intimately associated with the processes of deglaciation. An example of the former are basins
which are the result of glaciotectonic thrusting of bedrock and
drift during advance of the ice margin, common over large
areas of the northern Great Plains. Lakes can be created both
within the irregular ridge and furrow topography of the
deformed thrust blocks (see Aber, 1993,Fig. 13; Christiansen
and Whitaker, 1976, Fig. 3) and in source depressions up-ice
of the ridges (see Bluemle, 1991,Fig. 3 1-33). Another example of a lake basin related to active glacier ice is those lakes
occupying groundwater extrusion features (Boulton and
Caban, 1995), which may be quite common on the northern
Great Plains (Downey, 1969; Christiansen et al., 1982;
Bluemle, 1992,1993;Evans et al., in press). Such basins were
created when compressive flow near the ice margin resulted
in overpressuring of shallow groundwater aquifers, with consequent extrusion occurring proglacially. This process,
which produces generally circular basins with raised rims,
may account for some features previously attributed to
melt-out of stagnant ice (see below). The geolirnnology of
these basins is controlled, to a large degree, by the dynamics
of the continued groundwater flow and the composition of the
groundwater solution. For example, the groundwater aquifer
that supplies Howe Lake in southeastern Saskatchewan
(sandstone of the Cretaceous Mannville Group) contains
fresh water; the lake is therefore anomalously fresh despite its
closed basin and high evaporationlprecipitation ratio
(Christiansen et al., 1982; Last, 1991a). In the case of several
lakes in North Dakota, the groundwater aquifer contains
saline brines; these lakes are therefore anomalously saline, in
spite of a relatively humid climatic setting (Bluemle, 1992).
The vast majority of lakes on the northern Great Plains are
the product of deglacial landscape evolution. A few, such as
lakes Manitoba and Winnipeg in Manitoba (Fig. 3A) and the
Quill lakes complex in Saskatchewan, are direct ancestors of
large proglacial lakes formed by ponding of meltwater
against the retreating margin of the Laurentide Ice Sheet.
These modern remnants are important sources of late Pleistocene and early Holocene paleoenvironmental information.
Most lakes, however, resulted from melting of stagnant glacier ice and are not sedimentologically related to former
ice-contact lakes. Slow melting of ice buried beneath a thick
layer of superglacial debris created the variety of irregular
depressions and topographically closed drainages that characterizes much of the region (Fig. 3B; Gravenor and Kupsch,
1959; Clayton and Cherry, 1967; Clayton and Freers, 1967;
Christiansen, 1973). These basins tend to be small, but may

contain great thicknesses of sediment. For example, North


Ingebrigt lake in southwestern Saskatchewan, today a
hypersaline playa basin less than 1 krn2 in area, contains more
than 50 m of Holocene salt and clay. A final example of a
modem lake basin related to deglaciation is provided by those
lakes with an obvious fluvial origin, evidenced by their long,
linear, riverine morphologies (Fig. 3C). Most of these basins
represent scours or other depressions within glacial meltwater channels, some of which have been shown to occupy older
drainage valleys (Rueffel, 1968b).
Postglacial geomorphic activity, principally landsliding
and sand-dune migration, has created localized damming of
some drainages and the formation of generally shallow lakes.
Other lake basins have unusual origins. Christiansen (1971)
and Gendzwill and Hajnal (1971) demonstrated that Crater
lake, a small circular lake about 6 m deep located near
Yorkton, Saskatchewan, occupies a collapse chimney created
by dissolution of the Prairie Evaporite, some 900 m below the
surface.
Basin morphology exerts a profound influence on the
sedimentary processes and resulting spatial distribution of
detrital sediments within the lake (HAkanson, 1977; Timms,
1992; Talbot and Allen, 1996), The shape characteristics of
the basin similarly help control the distribution of chemical
precipitates in lakes (Jones and van Denburgh, 1966; Reeves,
1968; Eugster and Kelts, 1983). The lakes of the Great Plains
region exhibit a wide range in basin size, shape, and depth
(Hammer and Haynes, 1978; Mitchell and Prepas, 1990).In a
morphometric survey of lakes along a north-south (climatic)
transect through eastern Alberta, Campbell et al. (1994) and
Campbell (1997) demonstrated a strong sensitivity of lake
area and volume to a measure of the evaporationlprecipitation
ratio. Last and Schweyen (1983) and Last (1994a) classified
lakes of the northern Great Plains on the basis of size, depth,
and degree of permanence (Fig. 4) using an expanded database of about 500 prairie lakes. Most lakes in the northern
Great Plains are small and shallow (Fig. 3B), although Lake
Winnipeg (Fig. 3A) is the seventh largest lacustrine basin in
North America and the largest lake in western Canada located
off the Precambrian Shield. Lake Manitoba (Fig. 3A) is North
America's thirteenth largest lake and Canada's largest saline
lake. Of the lakes in southern Saskatchewan and Alberta, two
of the largest, Big Quill and Old Wives, are among the six
largest inland saline bodies of water on the continent.
Among the small lakes (<I00 km2), only about 8% have
mean depths greater than 3 m (Fig. 4). However, these small
deep basins are very important for paleoenvironmental
research because their sediment records often contain undisturbed, finely laminated sequences. These deposits are less
susceptible to wind and current redistribution than those in
shallow-water basins, and also less subject to diagenetic
changes resulting from subaerial exposure or subaqueous
chemical fluctuations (Richardson, 1969; Teller and Last,
1990). Several of these small deep basins are also
meromictic, which further enhances their appeal for
paleolimnological study (Schweyen, 1984; Last and
Schweyen, 1985; Last and Slezak, 1986).

GSC Bulletin 534

Of the many small shallow lakes, distinction can be made


based on the degree of permanence of the water body. The
term 'playa' has many definitions, synonyms, and, in some
cases, contradictory connotations (Reeves, 1968,1972; Neal,
1975). Previous researchers on the Great Plains (e.g. Lockhart and Last, 1984; Osterkamp and Wood, 1987; Gustavson
and Holliday, 1992) considered a playa to be any low, seasonally flooded but intermittently dry basinal area, and deemed
the term to be generally synonymous with ephemeral lake,

slough, or wetland. Rosen (1994) recommended playa be


restricted to continental basin settings characterized by an
annual net negative water balance and a capillary fringe close
enough to the surface that evaporation will cause groundwater to discharge to the surface. About half the lakes with
mean depths less than 3 m exhibit playa characteristics in that
they fill with water during the spring and summer and usually
dry completely by late summer. However, not all of these
shallow, ephemeral lakes exhibit groundwater discharge.

Figure 3. Examples showing the diversity in lake basin morphology in the northern Great Plains:
A) composite of satellite images show the large lakes of the northern Great Plains (lakes Winnipeg,
Manitoba, Dauphin, Winnipegosis, and Cedar), which are direct ancestors of giant proglacial lakes;
B) portion of a topographic contour map (left)from NTS area 72P/13, near Viscount, Saskatchewan, and
aerial photograph (right, NAPL A11069-296) from near North Battleford, Saskatchewan (Mollard and
Junes, 1984, Fig. 3-22) showing the high density of small lakes in areas of hummocb moraine; C) aerial
photographs of riverine lake basins in southern Saskatchewan (left, Verlo lake; NAPL A1 1069-298; centre,
North lngebrigt lake; NAPL A1 1099-80; right, Ceylon lake; NAPL 11099-82).

W.M. Last
-

- -

LAKES OF THE NORTHERN


GREAT PLAINS, WESTERN CANADA
(532 BASINS)

LARGE BASINS
(> 100 km2)

DEEP BASINS

SHALLOW BASINS

SMALL BASINS
(< 100 km2)

DEEP BASINS SHALLOW BASINS

Gallup (1978), and Lieffers and Shay (1983) in central


Saskatchewan; and Driver (1965) and Barica (1975) in the
Riding Mountain area of western Manitoba.
Chemical data exist for 5 11 lakes in the Canadian Great
Plains (Fig. 5), 10% of which are located in Manitoba, 72% in
Saskatchewan, and 18% in Alberta. Although most of these
data represent analyses of single samples, some are averages
of numerous samples collected over a period of months or
years. In general, the larger lakes (e.g. Winnipeg, Manitoba,
and Quill) have the longest temporal records, in some cases
dating back to the early twentieth century. However, no lake
on the Canadian Prairies has a continuous monitoring record
of more than four decades.

~ [ m
PERMANENT

INTERMITTENT

Salinity and composition


Figure 4. Summary of lake-basin morphology types in tlae
northern Great Plains (modified from Last, 1994a, Fig. 2).

WATER CHEMISTRY AND PROPERTIES


Ln this region, there are numerous ponds and small lakes in

the hollows among the hills, most of them being more or less
brackish or nauseous to the taste from the presence of
sulphates of magnesia and soda and other salts. During the
dry season of autumn, the water evaporates completely from
many of these ponds leaving their beds covered by the dry
white salts, which look like snow and are blown about in the
wind.

- Bell (1874)

Most of the millions of lakes on the Great Plains of western Canada are saline and, throughout much of the region,
ponded saline and hypersaline brines are the only surface
waters present. The importance of these lakes as waterfowl
nesting and staging grounds (Scott and Scott, 1986; Batt et al.,
1989), combined with water demands associated with future
agricultural, industrial, and urban development, will likely
lead to conflicts and potential environmental problems
(Cameron, 1986; Herrington et al., 1997).Thus, considerable
effort has been made to collect water composition data and
information relevant to the management of water resources.
The first analyses were reported from these lakes over 120
years ago (Bell, 1874). Data compilations can be found in
Moore (1939), Rawson and Moore (1944), Thomas (1959),
Rutherford (1970), Inland Waters Directorate (1975, 1980,
1984), Hammer (1978b), Last et al. (unpub. report, 1983),
Lambert (1989), Mitchell and Prepas (1990), and Last
(1991a.. b).
sources of information on water
, Other im~ortant
composition for p;airie lakes include Govett (1958),
Bierhuizen and Prepas (1985), and Evans and Prepas (1996)
in central and eastern Alberta; Hartland-Rowe (1966) in
southeasternAlberta; R6zkowski (1967) and R6ikowska and
R6ikowski (1969) in the Moose Mountain area of southern
Saskatchewan; Driver and Peden (1977), Schwartz and

Discussion of water salinity, although a very simple concept,


is often confusing due to the large variety of methods used to
measure this basic parameter (Appendix B) and a plethora of
related nomenclature (see reviews by Williams, 1967;
Carpenter, 1978; Hammer, 1986). Most biological limnologists use the following classification scheme: fresh water
(<l%o), subsaline (1-3%0), hyposaline (3-20%0), mesosaline
(20-50%0), and hypersaline (>50%0).Groundwater researchers refer to fresh water (<1%0), brackish water (1-10%0),
saline water (10-100%0), and brine (>100%0).Finally, most
geoscientific literature uses the following classification:
fresh water (c 3%0), saline (3-35%0), and hypersaline
(>35%0).
Lakes of the northern Great Plains range in salinity from
relatively dilute water (0.1 ppt total dissolved solids (TDS))
to brines that are more than an order of magnitude more concentrated than normal sea water (Fig. 6). Given the enormous
geographic area and the varying hydrological, geomorphic,
and climatic settings, it is not surprising that lake waters also
show a considerable range in ionic composition and concentration. Although it is misleading to generalize using means,
the 'average' lake water has about 40 ppt TDS and shows
Na>Mg>Ca>K and S04>C1>HC03>C03. Early investigators emphasized a strong predominance of Na and SO4 in the
lakes (Cole, 1926; Tornkins, 1954a). It was not until the
1980s that the compositional range and degree of diversity of
the lakes became evident. It is now apparent that, not only is
there a complete spectrum of salinities, but virtually every
water chemistry type is represented in lakes of the region
(Fig. 7). Nonetheless, sulphate- and carbonate-rich lakes
clearly dominate the Great Plains, accounting for over 95% of
the total. The paucity of C1-rich lakes makes the region
unusual compared with many other areas of the world (e.g.
Australia, western United States).
Most solutes in the lake waters increase in concentration
with increasing total salinity (Fig. 8). Sulphate, sodium, and
potassium ions show the most statistically significant correlations with TDS, whereas calcium and carbonate concentrations are less directly related to salinity. The proportions of
solutes also show a systematic change with salinity (Fig. 8).
Sulphate and sodium increase in relative ionic proportion

GSC Bulletin 534

ALBERTA

SASKATCHEWAN

MANITOBA

/I

MANITOBA

I/

Edmonton

... :

.%!I.

Lakes with
water chemistry
data

i!
L

ALBERTA

'

SASKATCHEWAN

Edmot~jon

,'

-2

'

\..
\L

Lakes with
surface
sediment data

,Calgary
I
I

200 km

.\

----

(
.

ALBERTA

i!
L

Edmonton

*I

-8

,'

ri

SASKATCHEWAN

~-

I.

.\,\ , .
-.:I

. ; . . ,>.,.'
.' :;....
.
. ..
.I .?'.'. .
'.

\,

Lakes with '\


c?.lg8ly
stratigraphic \',
"
,
data
I

'. -

.-

,,.

I...

..

I-

. I\

\ -

1.-

.'

1'.
1

\I
\

I,

' Regiria

'I'

/I

f'

\..

,,

. I

/I

//

! .

*a

MANITOBA

.I

200 km

'

\.

'

\I . .

,,>,

Figure 5. Locations of lakes with geolinznological data.

iI 1

\A!in~iipe:~
I

W.M. Last

from less than 20% equivalents in dilute lakes to generally


greater than 50% in lakes with more than 10 ppt TDS. Calcium and bicarbonate proportions show an inverse relationship with salinity, decreasing from greater than 60%
equivalents in the dilute waters to nearly 5 % in lakes with
more than 25 vpt TDS.
A

The relatively uniform distribution of lakes in the Great


Plains for which water chemistry data exist permits examination of spatial patterns of ionic concentration and proportion.
Such analyses have been undertaken by Winter (1977) and

Gorham et al. (1983) for sections of the northern Great Plains


in Minnesota and North and South Dakota, and by Last and
Schweyen (1983), Last (1988, 1992a), and Lambert (1989)
for the Canadian Prairies. Lakes with the highest Na+K, Mg,
and SO4 concentrations generally occur in east-central
Alberta and west-central to southern Saskatchewan, whereas
lakes with high HC03+C03 and C1 contents are found in central Alberta and western Saskatchewan.Lakes with relatively
low proportions of Ca and Mg occur in the northern and tentral parts of the Plains.

Figure 6.
Frequency distribution of average salinities for
511 lakes in the northern Great Plains. Mean
salinity for lakes of the region is 39.4 ppt TDS.
Salinity axis is loglo (modified from Last,
1992a, Fig. 2).

10

Salinity (ppt)

Figure 7. Ternary diagrams of average ionic composition (percent equivalents) of 511 lakes in the
northern Great Plains (modified from Last, 1992a, Fig. 5).

GSC Bulletin 534

I.

Anions: trends in concentratlon versus sallnlty

Cations: trendiln ionic proportidn versus salinity

0.1

10

100

Anions: trends in lonlc proportlon versus sallnlty

1000 0.1

10

100

1000

Total salinity (ppt)

Total salinity (ppt)

Figure 8. Ionic concentration (upper graphs) and ionic proportion (lower graphs) versus salinity for 511
lakes in the northern Great Plains. The lines in the upper graphs are second-order bestfit linesfor molality
versus salinity for each ion; the lines in the lower graphs are linear regressions of per cent equivalents
versus salinity for each ion.

A complicating factor in characterizing the geochemistry


of many lakes on the northern Great Plains is the strong seasonality of water levels (especially in playas), which gives
rise to dramatic changes in both ion concentrations and ratios
(Lieffers, 1981; Lockhart, 1984; Lockhart and Last, 1984;
Hammer and Parker, 1984; Last, 1984a, 1987, 1989b). For
example, Ceylon lake, a salt-dominated playa in southern
Saskatchewan, annually undergoes changes in concentration
from about 30 ppt to greater than 300 ppt TDS. This lake also
exhibits dramatic fluctuations in ionic ratios on a seasonal
basis, from a Na-(Mg)-SO4-.HCO3 type in early spring to a
Mg-(Na)-Cl-SO4 composltlon by fall. Hammer (1978b,
1986) summarized the short-term temporal changes in salinity and chemistry of several other saline lakes in the region.
Unfortunately, only a few basins in the northern Great Plains
have undergone periodic detailed sampling over a span of
several years.

Controllingfactors
The major ion composition and concentration of lakes on the
northern Great Plains is the product of 1) the complex interaction between precipitation, meltwater and unconsolidated
sediments and bedrock in the drainage basin; 2) the composition and amount of groundwater rechargeldischarge and
streamflow in each basin; and 3) a wide variety of other physical, chemical, and biological processes operating within the
water column itself. Several geochemical approaches have
been taken to help understand the major factors controlling
surface-water chemistry. These include mass balance calculations, thermodynamic equilibrium considerations, and statistical evaluations (Braitch, 1971; Hardie and Eugster, 1978;
Drever, 1988). In western Canada, both mass balance and
thermodynamic calculations have proven valuable in deciphering many of the intrinsic (i.e. within the drainage basin)
processes important in water composition on a local scale
(~6ikowski,-1967;Wallick and -Krouse, 1977; Wallick,

W.M.

1981; Last, 1984b; Slezak, 1989). In contrast, on a regional


scale, various statistical techniques have been successfully
applied to help understand the relationships between the
water composition and extrinsic environmental factors such
as climate, bedrock, geomorphology, and till composition
(e.g. Dean and Gorham 1976; Winter 1977; Last, 1992a). As
emphasized by Evans and Prepas (1996), these statistical
approaches lack the ability to resolve the often important
local conditions and processes. However, they are essential to
our overall understanding of the lacustrine geochemical setting of the region.
One of the most straightforward ways to analyze theinterrelationships within a data set is to examine the simple linear
correlations that exist among the various parameters. As
expected, the concentrations of Na, Ca, and Mg in the brines
all have significantly positive correlations, as do SO4 and C1.
In addition, the ion pairs Mg-SO4, Mg-Cl, and Na-Cl tend to
be strongly covariant. Importantly, the concentrations of Na
and SO4 do not show statistically significant linear correlation, suggesting that different suites of processes affect the
abundance of each of these ions.
By combining a variety of morphological (basin area,
maximum depth), geological (bedrock type, depth to bedrock, till type), hydrological (drainage basin area, number of
streams entering lake, elevation, groundwater composition),
and climatic (mean annual precipitation, evaporation, temperature) variables with the 500-lake water chemistry database, Last (1988, 1992a) and Lambert (1989) used R-mode
factor analysis to identify a set of seven statistical factors that
explained more than 90% of the variance in the data. Their
interpretation of these statistical factors is that the most
important controls on water composition and concentration
on a regional basis are composition of inflowing groundwater, evaporationlprecipitation,and elevation or position of
the basin within the drainage basin. Variables related to bedrock type, glacial drift composition, fluvial input, and lake
morphology are statistically less important.

Last

of the easternmost region, where hypersaline groundwaters


from the Paleozoic sequence are clearly important contributors to lake salinity (van Everdingen, 1971; Last, 1984b).
In contrast to the deep Paleozoic sequence, relatively
shallow Cretaceous and Tertiary bedrock has been implicated
as the source of at least some of the dissolved components in
the lakes (Cole, 1926; Sahinen, 1948; Wallick and Krouse,
1977; Wallick, 1981). Many researchers have also stressed
the close association of the more saline brines with buried
preglacial channels, and concluded that these buried valleys
act as conduits for groundwater supplying dissolved material
to the lakes (witkind, 1952; ~ueffel,1968a; Freeze, 1969).
Finally, there is considerable support for the source of the
ions being largely the Quaternary deposits within which the
lakes are situated (Rbzkowski, 1967; Rutherford, 1970;
Sloan, 1972; Moran et al., 1978; Timpson and Richardson,
1986; Arndt and Richardson, 1992; LaBaugh et al., 1996).
Although defining flow patterns and groundwater dynamics
in this setting is exceedingly complex (Winter, 1978, 1995),
Hendry and Schwarz (1982) suggested that, on a local basis,
glacial till is a large "soluble salt reservoir" which can proiide salinity to the shallow groundwater systems and thkrefore to the lakes. A variety of physicochemical and
biochemical reactions, including cation exchange, dissolution of feldspars and detrital carbonates, oxidation of
reduced-sulphur mineral species and organic matter, and precipitation of authigenic sulphate, carbonate, and silicate
phases in tills, have been documented which support this
hypothesis. In assuming a constant local (glacial drift) source
and unchanging source-water chemistry, Whiting (1974,
1977, 1986) and Doering (1982) used the level of salinity to
determine the amount of time required to accumulate the dissolved components in selected closed-basin lakes and
wetlands in the region, and to identify episodes at which the
now-closed basins overflowed.

MODERN SEDIMENTARY ENVIRONMENTS

Source of salts
The origins of the major ions in the lakes of the northern Great
Plains have been topics of considerable discussion in the scientific literature (see overview in Murphy (1996)). Early
research suggested that dissolution of deeply buried
Paleozoic evaporites could be a possible source for the dissolved components in the lakes (e.g. Grossman, 1968). The
occurrence of lacustrine basins and other geomorphic features whose origin may be ascribed to collapse of
salt-solution chimneys (Christiansen, 1967a, 1971) supports
this contention. However, Last and Schweyen (1983) and
Last and Slezak (1987a) maintained that dissolution of the
Devonian Prairie Evaporite could not contribute waters having ionic ratios compatible with the majority of prairie lakes.
Indeed, there is no consistent spatial correspondence between
the surface features and chemistries and the subsurface
Paleozoic rocks in most of the Great Plains (Slezak and Last,
1985; Last, 1988, 1989a, 1992a), with the notable exception

The study of sedimentary rocks, as compared with igneous


rocks, is somewhat hampered by the poverty of terminology.
- Moorhouse, 1959, p. 333

Few of the millions of lake basins in the northern Great


Plains have been examined from either a sedimentological or
stratigraphic perspective. Of the approximately 500 lakes for
which water chemistry has been documented, only about
20% also have data on sediments (Fig. 5). As with the early
work on brine composition, initial sedimentological efforts
stressed the dominance of sodium sulphate salts, and were
directed mainly toward basins with large reserves of economically important industrial minerals (Appendix A). It is now
apparent that lakes in the Canadian Great Plains exhibit a
complete spectrum of sediment types, from basins dominated
by allogenic (detrital) material to those in which relatively
pure evaporite minerals are forming.
The modern sediments in most lakes of the region comprise mixtures of coarse to finely crystalline salt and
organic-rich silty clay in the offshore portions of the basins,

GSC Bulletin 534

are sometimes collectively referred to as chemical sediment,


as opposed to clastic sediment which is dominated by the
allogenic fraction.

grading to somewhat coarser clastic material (sand and silt) in


the nearshore areas. Due to their small size and negligible
fetch, most lakes have only a narrow shoreline-nearshore
zone. However, in larger basins and lakes with long fetches
(e.g. riverine lakes), coarse clastic sediments can extend farther out into the basin (Veldman, 1969; Sack, 1993). Deltas
are rare except in the larger basins, such as lakes Manitoba
and Winnipeg, owing to the paucity of fluvial input. Away
from nearshore areas, however, there is usually little variation
in grain size in any of the basins (Icushnir, 1971; Whiting,
1977; Brunskill and Graham, 1979; Last, 198413, 1996~1,b;
Henderson and Last, 1998).

The allogenic component in most of the lakes of western


Canada consists of a mixture, in approximate decreasing
order of abundance, of clay minerals, quartz, carbonates,
feldspars, and fen-omagnesian minerals. The relative proportion of each of these detrital mineral groups is generally similar to that of the surrounding glacial debris. Less chemically
stable constituents are susceptible to loss by weathering processes in the drainage basin (Ehrlich and Rice, 1955; Mills
and Smith, 1971: Teller and Last, 1979). Another factor controlling the occurrence and distributidn of detrital components is the inherent size fractionation of various minerals.
F o r e x a m p l e , because t h e terminal g r a d e f o r glacier-comminuted dolomite is medium to coarse silt
(Dreimanis and Vagners, 1971), the modern sedimentary
facies of the larger lakes in the region show a gradation from
dolomite-enriched sediment near the shoreline to dolomite-depleted deposits offshore (Brunskill and Graham,
1979; Teller and Last, 1979; Last, 1982; Henderson and Last,
1998).

The organic content of the modern sediments is generally


moderate to low, but may be as much as 45% (by weight) of
the total sediment in some basins. The organic content of sediments from playa basins and shallow lakes is not significantly different from that of most of the perennial deep-water
basins. However, the chemical composition of this organic
matter can vary significantly. Playa lake sediments in the
region consistently show much lower Rock-EvalTMhydrogen
indices (HI) and somewhat higher oxygen indices (01) relative to both shallow and deep perennial lakes (Fig. 9),
reflecting a more terrestrial source for the organic material in
the playas. Similar results are evident for lakes in Minnesota
(Dean and Stuiver, 1993).
u

Although the sediments of only a small number of lakes


have been examined for detailed clay mineralogy, the most
common layered silicates are kaolinite, illite, smectite, and
chlorite (Kushnir, 1971; Egan, 1984; Last, 1984b, 1996b;
Last and Schweyen, 1985). In lakes west of the Manitoba
Lowland, sediments are almost completely dominated by
smectitic clays, reflecting the bentonitic Cretaceous shale
that underlies most of the area. The only detrital carbonate
minerals identified in the lakes are dolomite and calcite. In
bulk (i.e. non-size-fractionated) samples, CaMg(C03j2 is
usually considerably more abundant than CaC03, likely
reflecting the relative abundance of these minerals in the tills
and bedrock of the region, as well as the higher solubility of
calcite (Last, 1982, 1996a; Ghebre-Egziabhier and
St. Amaud, 1983a, b; Henderson and Last, 1998).

Detrital versus chemical sediments


Geoscientists recognize three basic types of sediment in most
lacustrine basins (Jones and Bowser, 1978): 1) allogenic or
detrital material derived from weathering and erosion of soils
and bedrock of the watershed and transported to the lake by
fluvial, gravity, or eolian processes; 2) endogenic sediment
originating from processes occurring entirely within the
water column of the lake; and 3) authigenic or diagenetic
material resulting from processes that occurred once the sediment was deposited. The endogenic and authigenic fractions

FF

100

200

Figure 9.

Elk

300

400

500

Oxygen index (mg C02/g OC)

600

700

Rock-EvalT" hydrogen indices ( H f )and oxygen indices ( O f )for modern s~~rJCicia1


sediment sanzples from
lakes iiz western Canada and elsewhere. Indices
express concentrations per graiiz of orgarzic carbon
(OC). The grey shaded area encompasses 90% of the
50playa lakes. The vertically hatched urea shocvs the
range of indicesfrom deep-water maar lakes in southeastern Australia; the diagonally hatched area shows
the range of indices froin the Aral Sea nzodern
sediiizents. F F , Freefight Lake, southwestern
Saskatchewan; Din, Deadmoose Lake, south-central
Saskatchecvaiz; L m , L i t t l e M a n i t o u Lalce,
sontlz-centrcil Saskatchewan; Wald, Waldsea Luke,
south-central Saskatchewan; Red, Redberry Lake,
central Saskatclzewarz; Man, Lake Manitoba,
Manitoba; Elk, Elk Lalce, Minnesota (Dean and
Stuiver, 1993).

W.M. Last

More than 50 nondetrital minerals have been identified


from the lakes of the northern Great Plains (see compilations
in Last (1989a), Shang and Last (1999)), reflecting the wide
range of water compositions present. Two main types of
endogenic precipitates occur in modern sediments: very soluble salts, comprising mainly sodium and magnesium
sulphates and carbonates, and sparingly soluble precipitates,
including mainly carbonates, sulphates, and silicates (Last,
1984a). These can be further subdivided, on the basis of genesis and occurrence within the lake (Fig. 10, l l ) , into

1) surficial efflorescent crusts and hardgrounds, usually


occupying nearshore and seasonally flooded areas; 2) massive and bedded precipitates, most often found blanketing the
floor of the basins from shallow marginal zones down to deep
central offshore areas; and 3) accumulations of salts and precipitates associated with either subaqueous or subaerial
springs.
Distinguishing endogenic from authigenic lacustrine
minerals is often exceedingly difficult (e.g. Talbot and Kelts,
1986; De Deckker and Last, 1989), requiring detailed

Figure 10. Examples of modem endogenic precipitates from lakes in the northern Great Plains: A) epsomite
(bfgS04.7H20) from Syboz~tsLake, southern Saskatchewan; GSC 1999-045A; B) dog-tooth mirabilite
(NazS04.1OH20) below an upper layer of bladed bloedite (NazMg(S04)2-4H20)
from Ceylon lake, southern
Saskatchewan; GSC 1999-0458; C) hopper-shaped mirabilite from Ceylon lake showing incl~isionsof brine
shrimp; GSC 1999-045C; D) lenticular dendritic crystals of epsomite and mirabilite covering the floor of
Chain lake, southwestern Saskatchewan; GSC 1999-0450; E)floating rafts and crusts ofthenardite (NazS04),
Vincent Lake, southwestern Saskatchewan; GSC 1999-045E; F) accumulations of mainly bloedite forming
concentric rims around spring openings in Ceylon lake; GSC 1999-0451;; G) massive aggregate of halite
(NaC1) from Bitter Lake, southwestern Saskatchewan; GSC 1999-0456; H) dendritic acci~mulationof
acicular glauberite (NazCa(S04)2)from North lngebrigt lake, southwestern Saskatchewan; GSC 1999-045H.
Photographs by W.M. Last.

GSC Bulletin 534

and Hsu, 1978; Dean, 1981; Eugster and Kelts, 1983; Dean
and Fouch, 1983). Last and Schweyen (1983) found that
nearly all of the 131 lakes they surveyed in the northern Great
Plains were strongly supersaturated with respect to these carbonate minerals. In most of these lakes, supersaturation was
due to the seasonal uptake of C 0 2 by primary organic productivity. The specific carbonate mineral precipitated from the
supersaturated solution is controlled, to a major degree, by
the cations in solution (in particular the ratio of magnesium to
calcium in the water; Moller and Kubank (1976), Last
(1982)). The high Mg/Ca ratios that characterize the lake
waters of the Great Plains give rise to a carbonate mineral
assemblage dominated by aragonite (the orthorhombic form
of CaC03) and Mg-bearing carbonates, such as dolomite,
magnesite, huntite, and Mg-calcite.
The discovery of nondetrital dolomite in the lakes of the
Great Plains has been an important contribution to our understanding of the genesis of this economically important mineral. Dolomite formation and dolomitization in the
sedimentary realm are subjects of long-standing interest and
study (Chilingar et al., 1979; Last, 1990). The complex 'dolomite problem' has been summarized in many reviews
(e.g. Zenger and Mazzullo, 1982; Shukla and Baker, 1988;
Purser et al., 1994). In its simplest form, the dolomite problem is the enigma that the mineral dolomite (CaMg(CO&) is
a very common component of ancient rocks, but is very rare
in Holocene sediment, does not occur as a primary precipitate
from marine water of normal salinity and composition, and
cannot be readily synthesized in the laboratory at low temperatures and pressures.

Figure 11. Primary and early (penecontemporaneous)


dolomite at Freefight Lake, southwestern Saskatchewan:
A) lithified crust on exposed mudflat. Photograph by
W.M.Last; GSC 1999-0451; B) SEM photomicrograph of
poorly crystalline modern dolomite from the deep-water
offshore sediments.

petrographic and geochemical studies. To date, few lakes and


lacustrine sequences in western Canada have been examined
in such detail. Although a complete discussion of the genesis
(viz. endogenic versus authigenic) of the minerals identified
in these lakes is beyond the scope of this paper, some of the
more common or noteworthy occurrences will be summarized briefly.
Among the many endogenic and authigenic carbonates
identified in these lakes, aragonite, magnesian calcite, and
dolomite are the most common. Thermodynamic
supersaturationand inorganic precipitation of carbonate minerals are common in lakes of all types on a global basis (Kelts

The occurrence of penecontemporaneous (i.e. very early


diagenetic) dolomite in the surficial offshore sediments of
Devils Lake, a 50 km2 perennial saline lake in northeastern
North Dakota (Callender, 1968; Owen at al., 1973) was one of
the earliest documented examples of lacustrine dolomite formation in the world. The Devils Lake dolomite demonstrated,
for the first time, that dolomitization could take place in solutions of moderate salinity (approx. 20 ppt TDS) and in water
with high sulphate content. Modern primary and early
diagenetic dolomite also occurs in numerous lakes in the
Canadian Great Plains (Last, 1990), including Waldsea,
Deadmoose, and Freefight lakes in Saskatchewan. Like that
of Devils Lake, the dolomite in both Waldsea and
Deadmoose lakes is forming subaqueously as a primary inorganic precipitate in sulphate-rich waters (Schweyen and Last,
1983; Schweyen, 1984;Boden, 1985; Last and Slezak, 1986).
In Freefight Lake, calcian (i.e. Ca-enriched) dolomite occurs
in the deep-water offshore sediments and also in the
subaerially exposed mudflats surrounding the basin (Last and
Slezak, 1987b; Slezak, 1989; Last, 1993a). Within the mudflat sediments, it is often associated with aragonite and
Mg-calcite, and may be an early diagenetic product forming
in response to the strong evaporative flux experienced during
the ice-free season (Fig. 11A). In contrast, in the organic-rich,
highly reducing offshore sediments, dolomite is associated
with a variety of soluble sulphate salts and pyrite, and is forming at the sediment-water interface as a primary precipitate in
response to the increased alkalinity brought about by sulphate
reduction in the anoxic monirnolimnion of the lake (Fig. 11B).

W.M. Last

The mineralogy of the endogenic and authigenic


sulphates in Great Plains lakes is complex. Not only is the stable sulphate mineral suite controlled by the ionic composition
and cation ratios in the brine, but the temperature at which the
precipitation occurs is also very important (Fig. 12). The
effect of temperature is most obvious in the playas (Last,
1984a, 1987, 1989b) and shallow lakeslwetlands (Timpson
and Richardson, 1986; Timpson et al., 1986; Skarie et al.,
1987) where the annual cycle of sediment precipitation and
dissolution is readily apparent. However, deep perennial salt
lakes are also strongly influenced by annual temperature
regimes. For example, surface waters in Freefight Lake
become supersaturated with respect to mirabilite
(Na2S04-10H20) and epsornite (MgSO4-7H2O) during the
winter due to low temperatures of the upper rnixolirnnion (-10
to -15C depending on salinity) and the formation of an ice
cover, which increases the brine concentration in the upper
metre. The precipitated crystals undergo corrosion and dksolution as they settle through the undersaturated rnixolimnion
(-5 to 5C) below the ice cover. Dissolution of these sulphate
salts is a strongly endothermic reaction, of sufficient magnitude to affect the temperature of the water (dissolution of
1 mol of rnirabilite absorbs more than 4 MJ of heat). This heat
must be acquired from the surrounding water, thereby lowering its temperature. However, sulphate salt precipitation
begins anew at the chemocline, despite a still higher temperature, due to the greater salinity. Gypsum shows a similar pattern of precipitation in the surficial waters, dissolution within
the rnixolirnnion, and precipitation through the monimolirnnion, except that the supersaturation at the surface is usually brought about during the ice-free season due to
evaporative concentration ofthe brine. Thus, the petrography
of the resulting deep-water sulphate salts in this basin records
a complex series of multiple precipitation and dissolution
events that are controlled largely by seasonal temperature
fluctuations within the water column.

Sedimentary processes and facies


The lakes of the northern Great Plains offer an excellent
opportunity to examine the processes of lacustrine sedimentation on both local and regional scales. The broad ranges discussed previously for each of basin morphology, hydrology,
degree of permanence, brine salinity, ionic composition,
sediment character, and composition result in a similar continuum of sedimentary regimes. In an effort to subdivide this
continuum, 'end member' types of lakes can be recognized. A
fundamental sedimentological distinction must be made
between those basins whose brines are shallow enough to permit periodic drying and those whose brines are deep enough
to maintain a relatively permanent water body.
The most significant processes operating in shallow intermittent basins include cyclic flooding and desiccation of the
playa surface; formation of salt crusts, efflorescent crusts,
hardgrounds, spring deposits, and intrasedimentary salts; formation of solution pits and chimneys; and periodic detrital
sedimentation by sheet flow and wind (Fig. 13). Although
comprehensive summaries of these basic playa processes can
be found in Eugster and Hardie (1978), Handford (1981),
Lowenstein and Hardie (1985), Smoot and Lowenstein
(1991), and Rosen (1994), most of these overviews deal with
lacustrine systems in warm and dry climatic regimes. The
five to six months of subfreezing temperatures experienced in
western Canada, as well as the high seasonal runoffs associated with snowmelt, dictate that some modifications and additional processes be considered when studying the playas of
the northern Great Plains (Last, 1987, 1989a, 1994a).
Despite the great range of sediment types and hydrological settings, the chemical and physical processes operating in
playa basins of western Canada create a discrete suite of modem sedimentary facies, although significant variations in the
development of these facies occur from basin to basin. An

Figure 12.
Solubility versus temperature of commonly
occurring endogenic precipitates in lakes of
western Canada. The vertical scale on the left
side of the diagram is relative to gypsum and
anhydrite.

10

20

30

Temperature (EC)

40

50

GSC Bulletin 534

outer shoreline-nearshore complex, comprising colluvium,


mud flat-sand flat, and beach facies, grades basinward into a
pan complex. The colluvium facies consists of a chaotic mixture of coarse to fine detrital material derived from the adjacent glacial deposits by mass wasting and creep. The mud
flats and sand flats are characterized by a mixture of sandy to
silty detrital sediments often capped by efflorescent crusts
(Fig. 14) and occasionally by hardgrounds. Unlike efflorescent crusts on mud flats in playas elsewhere (e.g. Australia,
southwestern United States; Hardie et al. (1978), Goudie and
Cook (1984)), these crusts are not monomineralic, but rather
usually consist of a complex mineral assemblage of both
hydrated and anhydrous Na, Na+Mg, and Ca sulphate and
sulphate-chloride salts (Keller et al., 1986a, b; Last, 1989a).
Carbonate hardgrounds and crusts are also present in some
basins. The mud flats-sand flats are areas of active shoreface
and beach processes during seasonally high-water levels, but
evaporation is the dominant process for much of the year.
Intense evaporation during the warm, dry summer months
results in upward capillary movement of shallow groundwater, ionic concentration, and intrasedimentary and crust

EVOLUTIONARY
CYCLE

Freeze-out
Precipitation

Fresh water
Influx

--:
ANNUAL

-+- - 3
-

The salt-pan complex usually constitutes more than 50%


of the playa basin. Depending on weather and drainage-basin
conditions, it can be covered by brine for much of the ice-free
period, but is often exposed by late summer or early fall.
Overall, the facies is usually characterized by high endogenic
to detrital sediment ratios, although a complete gradation
exists from salt-dominated to mud-dominated pans. The most
distinctive feature of the pan facies is its annual cycle (Lockhart, 1984; Last, 1984a, 1989b). During spring, relatively
dilute inflow from melting snow and rainfall can dissolve
much of the very soluble salt that was precipitated in the pan
during the previous dry episode. This results in a rapid
increase in water salinity and dramatic changes in ion ratios.
These brine composi~onalchanges drive many of the

FRESH

EXTINCT

f.

mineral precipitation. Once efflorescent crusts have formed,


further evaporation is hindered and the mud flats can maintain a high moisture content throughout the dry season. In
some basins, the mud flats are colonized by extensive areas of
vegetation and cyanobacterial mats (Fig. 15). Sediments in
these areas are distinctively laminated (Last and Vance,
1997), organic rich, and often sites of intense carbonate rninera1 genesis and diagenesis.

Desiccation

CYCLE

'L
ssolutlon

.-

Evaporation &
Precipitation
Efflorescence

Precipitation

Cooling

Solar

DIURNAL Heating
CYCLE

Dissolution

Figure 13. Schematic illustration showing the three major


cycles (diurnal,annual, evolutionary) common to many of the
playas in the northern Great Plains. The long-term
evolutionary cycle shown is one of nine such cycles
recognized in these lakes. Photographs by W.M. Last.
GSC 1999-045P, -045R, -045Q

Figure 14. A ) Efflorescent crust developed on the


mudflat at Corral Lake, southeastern Saskatchewan;
GSC 1999-0455; B ) The crusts form seasonally and
the mineralogy is complex, consisting of a mixture of
thenardite, rnirabilite, konyaite (Na2Mg(S04)2.5H20),
and loeweite [Nal2Mg,(SO4),,~l5H2O); GSC 1999-045K.
Plzotographs by W.M. Last.

W.M. Last

penecontemporaneous diagenetic reactions (Egan, 1984;


Last, 1992b). Throughout the summer, brine salinities
continue to increase due to evaporative concentration. Even
relatively minor diurnal temperature changes will cause massive salt precipitation and dissolution. If brine remains in the
basin after the onset of freezing conditions, considerable
thicknesses of salt can occur due to freeze-out precipitation.

*A

.
't,

p;'

&

'
'
I

Seveial subfacies can be recognized in most playa pans on


the basis of mineralogy and crystal morphology (Last,
1989b). A zone of large dog-tooth rmrabilite crystals develops (see Fig. 10B) near the margins where pan sediments
interfinger with the mud flat-sand flat facies. These crystals
grow displacively downward into the soft, water-saturated,
mud-flat sediments. Further into the basin, the floor of the pan
usually consists of a mosaic of large, interlocking, bladed
crystals (see Figs. 10D, 10H). The specific sulphate andlor
carbonate mineralogy is controlled by the cation content of
the brine and by diurnal and seasonal temperature fluctuations. Warm temperatures favour species such as thenardite,
hexahydrite, anhydr~te,and thermonatrite; cooler temperatures encourage mirabilite, epsomite, natron, and gypsum
(see Fig. 10A).
Mineral deposits associated with spring openings can be
superposed on salt pan evaporitic or elastic sediments. Spring
openings are often sites of very rapid and massive salt mineral
precipitation (see Fig. lOF), because these are areas where
waters of different temperature and composition mix. The
springs frequently continue to discharge onto the playa floor
to form large low mounds and ridges after the pan has been
completely desiccated (see Cole, 1926, 1930) for descriptions of these large salt and mud mounds).
In contrast to the complex assortment of interrelated sedimentary mechanisms operating in playa basins, deposition in
perennial lakes is controlled by a smaller array of important
processes (Brunskill and Graham, 1979; Last, 1994a). Most
of these processes, such as shoreline erosion and deposition,
wave and current distribution of sediments, river inflow and
deltaic sedimentation, pelagic fallout sedimentation, and
slumping, turbidity flow, and interflow, are common to any
aqueous medium. However, several processes are either
unique to the perennial basins of western Canada or are of
such fundamental importance that they will be briefly discussed here.

Figure 15. Exanzples ofpustular and laminated nzicrobialites


from lakes of western Canada: A) surface of litlzified
microbialite (thronzbolite) from Dana Salt Lake, central
Saskatchewan. Photograph by W.M. Last; GSC 1999-045L;
B) cross-section showing faintly laminated algal bedding
from Freefight Lake, southwestern Saskatchewan (modern
su$ace of the microbialite is at the top of the photograph;
modifiedfrom Last and Vance (1997,Fig. 5)).Photograph by
W.M. h s t ; GSC 1999-045M; C)photo- micrograph offinely
laminated dolomitic microbialite-algal encrustation from
Muskiki Lake, central Saskatchewan.

One of the most critical processes is development of a


stratified water column. The influence of seasonal temperature stratification on carbonate mineral saturation and eauilibria in these lakes is well known (Last and Schweyen, 19'83,
1985; Last, 1984b). The superposition of thermal stratification on an already chemically stratified (meromictic) water
column greatly complicates mineral precipitationldissolution
reactions. This complexity means that interpretations of the
associated stratigraphic record are difficult, because even
subtle chemical andlor temperature changes can result in substantial changes in mineral precipitation and preservation.
For example, both evaporitic and biogenically induced carbonate precipitation occur in the upper, well mixed
epilimnion of Deadmoose Lake. Much of this carbonate
material, however, is dissolved upon passing through the

GSC Bulletin 534

thermocline into the anoxic, lower pH water of the


hypolimnion, so modern carbonate-rich muds are found only
at water depths of less than 8 m (Boden, 1985; Last and
Slezak, 1986). In contrast, gypsum precipitation occurs in the
water column below a depth of about 12 m and is preserved in
the sediments of the monimolimnion. Although the surface
waters become supersaturated with respect to CaS04-2H20
during the winter and gypsum precipitation occurs, sediment-trap data indicate that this gypsum is redissolved before
reaching the chemocline (Last, 1994b). Similarly, mirabilite
precipitation occurs at the chemocline and in the surface
water during winter. However, it is only preserved in the sediments of the monimolimnion at water depths of greater than
20 m.
Another important process in many of the perennial lakes
of the Great Plains is the generation and accumulation of
deep-water soluble salts (Fig. 16). Although many of these
lakes are characterized either by mixed endogenic-allogenic
sediments or by entirely allogenic deposits (Last and
Schweyen, 1983), it has been shown that, in some basins, soluble salts and sparingly soluble minerals, including clays, are
forming in deep-water offshore environments (Boden, 1985;
Last and Slezak, 1987b; Slezak, 1989; Sack, 1993; Last,
1993b, 1994b; Sack and Last, 1994). Sedimentologists have
long recognized that there are very few examples of modern
sedimentary environments in which deep-water evaporite
mineral formation is occurring. This paucity is problematic
because many ancient evaporitic sequences have been interpreted as forming in deep water (Kendall, 1984; Peryt, 1987;
Warren, 1989). Thus, the perennial lakes on the Great Plains
in which deep-water salts are forming today, including
Deadmoose, Little Manitou, and Freefight lakes, provide a
critical analogue in helping to understand the sedimentology,
geochemistry, and stratigraphy of these ancient deposits.
All of the lakes noted above are dominated by sodium,
magnesium, and sulphate ions, but the mineralogy of the
modem offshore precipitates is somewhat different in each
lake and is controlled by the specific ionic ratios in the brines.
The mechanisms by which these salts form also vary. The
diurnal and seasonal temperature fluctuations of the
mixolimnions in Freefight and Deadmoose lakes and the elevated salinities of their monimolimnions lead to a complex
multisite source for the precipitates within each basin. In Little Manitou, which is considerably shallower than Freefight
or Deadmoose, it appears that precipitation is initiated by
brine mixing, associated with both subaqueous spring discharge and a periodic breakdown of the chemically stratified
water column.
Although the mechanism of precipitation in all three basins appears to be temperature driven, in fact crystallization
ulti&ely depends only upon evaporation. ~ i n e r a genesis
l
will continue as long as the climatic-hydrological regime in
each basin maintains the present ratio of inflowing water to
evaporation. Any increase in the hydrological budget of these
lakes will terminate mineral formation. If, however, the current hydrological budget is maintained for some time
(approx. 500 years in the case of Freefight Lake) precipitation
will also cease due to shallowing of the basin and breakdown
of meromixis. A second point is that, once thermodynamic

Figure 16. Examples of deep water evaporites: A) mirabilite


from Freefight Lake, southwestern Saskatchewan.
Photograph by W.M.Last; GSC1999-045N B) epsomite and
deep water endogenic/authigenic clays from Little Manitou
Lake, south-central Saskatchewan. Photograph by
W.M. Last; GSC 1999-0450; C) an iron-bearing smectite
from Freefight Lake.

W.M.

saturation (or supersaturation) has been reached by evaporative concentration of the brines in these basins, the rates of
mineral formation and deposition, as well as the relative
abundances of the various salts, are controlled by the differences in solubilities of the minerals within the range of temperatures experienced by the brine. In other words, the
amount of salt precipitated (i.e. the sedimentationrate) is controlled by the temperature differences rather than by the specific mineral solubility.
The rate at which these deep-water salts are accumulating
is extraordinary. Sedimentation rates at the sediment-water
interface in the deepest part of Freefight Lake have averaged
nearly 30 kg-m-2.a-' over the past 13 years. Although it must
be emphasized that these data are determined by sediment
traps and therefore do not represent 'net' accumulation, the
stratigraphic sequence recovered in the offshore areas of the
basin nonetheless suggests linear accumulation rates in
excess of 2 cm-a-l.Although such high rates of chemical sedimentation should be expected in an evaporitic regime
(Holser, 1979; Warren, 1989; Einsele, 1992), they were not
adequately documented in modern deep-water environments
until the investigation of these salt lakes in western Canada.

PALEO1,IMNOLOGY: POWER TOOLS


FOR THE NEW MILLENNIUM
It has long been my feeling that when a geologist gets into
trouble, he changes the climate.
- Beaty, 1971

The great diversity in morphology, composition, and


depositional processes found in the lakes of the northern
Great Plains offers an outstanding opportunity to study past
environmental conditions. Our present knowledge of the
postglacial sequences in these basins is still limited, but has
improved substantially over the past decade. Of the approximately 100 Canadian lakes for which modern
sedimentological data are available, the stratigraphic records
of about fifty have been investigated in any type of detail (see
Fig. 5). Ten years ago, there was less than a third of this number (cf. Last and Slezak, 1988). Clearly, interest in the
paleoenvironmental records of these lakes is high and growing (see reviews by Teller and Last (1990), Last (1992b,
1994a), Vance and Last (1994), Vance et al. (1995), Fritz
(1996), Sauchyn (1997), Last and Gosselin (1997)). Rather
than provide another overview of this rapidly evolving discipline, my intent here is to highlight several of the majorphysical, mineralogical, and geochemical tools that will continue
to help advance our understanding of the paleoenvironmental
records preserved in these lakes.

Texture
The texture of lacustrine sediments has been a mainstay for
paleolirnnologists for more than a century, and will continue
to provide important information about water depths and
sediment distribution in clastic-dominated sequences in the
Great Plains. AIlogenic particle size is generally used to interpret water depth by deducing the level of energy being

Last

imparted to the sediment-water interface at the sampling site.


Coarse-grained sediments (sand, gravel) are generally interpreted as having been deposited under relatively shallow-water, high-energy conditions (e.g. a beach, shoal, or
delta), whereas very fine-grained sediment (clay) requires
settling through a deep, essentially motionless water mass.
The textural maturity of the sediment, usually evaluated by
how well sorted the particle sizes are, is also useful for interpreting the effectiveness (consistency) of the mechanical
energy at the depositional site.
Because grain size is such a basic descriptive sedimentological parameter, examples of its use in paleolirnnology of
the northern Great Plains abound. Nearly all of the lacustrine
stratigraphic sequences examined in the region have recorded
grain size, although at considerably different vertical resolutions and degrees of precision. To cite just two well studied
examples, Teller and Last (1979, 1981) used a general
upward increase in mean grain size in the 14 m thick offshore
record of Lake Manitoba to indicate an overall shallowing of
the basin as this lake progressed from part of a giant, deep
(200 m) proglacial lake 12 000 years ago to a large but shallow (4 m) remnant today. Periodic incursions of the Assiniboine River and delta formation were recognized by the
occurrence of relatively coarse, well sorted clastic sediments
superposed on this overall coarsening upward trend. In a
detailed analysis of grain-size fluctuations in a 3.6 m long
core (approx. 4000 years) from Pine Lake in central Alberta,
Campbell et al. (1998) used the median size as a proxy for
past moisture conditions, coarse being interpreted as warm
and dry, and fine as cool and moist. Through the use of Fourier and nonlinear regression analysis, they recognized several major features, including the Little Ice Age and the
Medieval Warm Period, and identified a dominant 1500-year
cycle as well as several weaker century- to millenium-scale
cycles which correspond well to Milankovitch periodicity.
Despite the potential value of textural data, it must be
emphasized that interpretations are not straightforward.
Indeed, the basic relationship between grain size and water
depth must be used with caution. As stressed by Sly (1978),
Hdkanson and Jansson (1983), and Teller and Last (1990),
many other factors can interfere with this simple grain
size-water depth-energy relationship, including flocculation
of fine-grained material, basin morphology, weathering and
erosion characteristics of the watershed, and vegetation cover.

Structure
Laminated lake sediments are a great, largely untapped,
wealth of information for paleolimnologists working in the
southern Canadian Prairies. Last and Vance (1997) recently
summarized the range of bedding and lamination types found
in these lakes and provided several examples of how various
types of mineralogical, textural, and geochemical analyses
can provide information about past water levels, brine compositions, ionic ratios, hydrology, and water-column
stratification.
Although most of the extant lake basins cored to date
exhibit some degree of bedding and lamination, only a few
such sequences have been studied. One example is Chappice

GSC Bulletin 534

Lake, a small, hypersaline, groundwater-fed playa in southeastern Alberta whose modern brine is dominated by Na and
SO4 (Vance, 1991; Vance et al., 1992, 1993; Birks and
Remenda, 1999). Although endogenic carbonates are not
forming in the lake today, the 7000-year long stratigraphic
record preserved in the basin contains thick sequences of
finely laminated carbonate sediments (Fig. 17). The carbonate crystals and crystal aggregates making up these laminae
contain no petrographic evidence of reworlung, abrasion,
corrosion, or diagenetic alteration, suggesting the laminae
were generated by inorganic precipitation from within the
water column. The lack of detrital grains in these laminae and
the absence of rhythmicity indicate relatively rapid and
nonannual precipitation events. It is clear from the detailed
mineralogical composition of the sequence that the brine
underwent striking compositional changes in both carbonate
alkalinity and cation ratios.

Of interest to researchers studying the stratigraphic


records of more saline lakes is the occurrence of thin beds and
laminae of fine-grained detrital material, which commonly
occur interbedded with the evaporites. Although their hydrological significance and sedimentological interpretation are,
at present, enigmatic, these allogenic interbeds offer great
potential for supplementing our paleolirnnological interpretations in these salt-dominated basins. For example, if these
thin beds represent periods of increased influx of clastic
material from the surrounding watershed, this would imply
increased runoff and probably relatively deep and
nonhypersaline conditions (i.e. more humid conditions).
Conversely, the fine silt and clay laminae could be residual
products, the end result of dissolution and downward leaching of the salts due to prolonged desiccation of the playa and a
lowered groundwater table brought about by a more arid climate. An increase in aridity could also be interpreted from
these fine clastic sediments if they were deposited during
periods of increased eolian activity.
A promising, but as yet untested, avenue of future
research on Canadian Plains lakes is that of high-resolution
textural analysis of finely laminated sediments. Laser particlesize analysis (Agrawal et al., 1991), now allows rapid and statistically valid analysis of the complete size spectrum of individual submillimetre laminae from small-diameter cores.
This type of 'event' stratigraphy has been used successfully
on much thicker varves and rhythmically bedded sediments
of glacial and proglacial lakes in central and western North
America for many years, and should be applied to the finely
laminated Holocene sequences in the Prairies.

Mineralogy
In lacustrine systems dominated by inorganic endogenic mineral formation, as are most of the lakes of the western
Canadian Plains, composition of the inorganic fraction is a
leading paleoenvironmental parameter. The mineralogy and
geochemistry of the precipitated material is one of the most
d i r e c t and u n a m b i g u o u s i n d i c a t o r s a v a i l a b l e t o
paleolimnologists of past chemical composition of the lake
water. As pointed out by Shang andLast (1999), the theoretical basis for, and our knowledge of, the thermodynamics of
the reactions responsible for most of these minerals have
improved dramatically in the past few decades. Explicit
chemical paleoreconstructions are now possible for saline
lake waters once the detailed equilibrium mineral assemblage
is known. The importance of collecting mineralogical data
versus bulk sample elemental geochemical data, in order to
quantitatively reconstruct brine composition, cannot be overemphasized. Knowing that a sample contains quartz,
plagioclase, hexahydrite, and halite provides essential detail
unavailable through chemical analysis, which identifies only
the analytical proportions of Si, Na, Ca, Mg, Al, S, and C1.

Figure 17. Detailed carbonate mineralogy of individ~ialbeds


and laminae from Chappice lake, southeastern Alberta: A,
aragonite; C, calcite; D, dolomite; HMC, Mg-calcite; HM,
hydromagesite; PD, protodolomite; PHM, pseudohydromagnesite (modified from Last and Vance, 1997,
Fig. 12). Photograph by R.E. Vance.

Because salt minerals are thermodynamically and


kinetically responsive to even relatively minor changes in
brine composition (Braitsch, 1971; Sonnenfeld, 1984;
Sonnenfeld and Perthuisot, 1989; Renaut and Long, 1989),
basins that are most suitable for interpreting long-term chemical changes are those with relatively thick, continuous

W.M. Last

sequences of evaporites. Using changes in major-ion ratios


deduced from the evaporite mineralogy in 24 hypersaline
lakes, Last (1995) developed a complex series of nine different evolutionary sequences present in lakes of the Great
Plains. The most commonly occurring cyclicity in these lakes
was that of C03 * C03-SO4 * SO4 for the anions and Ca *
Ca-Mg Na * Na-Mg-Ca for the cations, although more
than 20% of the lakes examined exhibited no statistically significant cyclicity. Although lacustrine sequences dominated
by soluble evaporites are most suitable for these kinds of
brine reconstruction efforts, valuable information about past
lake-water compositions can be derived even from the
endogenic minerals of marginally saline and fresh-water systems. Carbonate minerals are among the most common
endogenic constituents in these lakes. Many studies, both in
the Great Plains and elsewhere, have documented the suitability of calcite and aragonite in paleolirnnology (Kelts and
Hsu, 1978; Dean, 1981; Eugster and Kelts, 1983; Dean and
Fouch, 1983; Dean and Megard, 1993;Haskell et al., 1996).

Although literature on the paleochemistry of lakes in the


Canadian portion of the Great Plains is discouragingly small,
it is clear that these salt lakes will provide additional information on the principles of brine evolution. Based mainly on
observational data provided by Miiller et al. (1972), it is generally held that the mineral sequence low-Mg calcite *
high-Mg calcite * aragonite * dolomite * magnesitel
huntite reflects increasing MgICa ionic ratios in the precipitating solution and, very likely, increasing salinities. Thus, it
has become routine to use the stratigraphic variation of these
endogenic species in the lacustrine basins of the Great Plains
to deduce past MgICa ionic ratios and salinity of the lake
water (Last, 1982; Last and Schweyen, 1985; Last, 1996c;
Vance and Last, 1996;Last and Vance, 1996).These mineralogical data can be particularly useful when interpreted in conjunction with stable isotopic analyses of either the inorganic
carbonates or biological components (Boden, 1985; Last and
Slezak, 1986; Van Stempvoort et al., 1993; Last et al., 1994;
Padden, 1996; Vance et al., 1997).
In addition to changes in size and sorting of detrital grains,
stratigraphic fluctuations in the relative abundance of the
minerals making up the allogenic fraction provide important
information about past lake-basin hydrology, source areas,
transport mechanisms, weathering processes, and erosion
characteristics. Last and Sauchyn (1993), Teller and Last
(1981), and Henderson and Last (1998) used the ratios of total
feldsparslquartz, plagioclaselK-feldspars, and calciteldolomite as qualitative indicators of chemical-weathering intensity in the associated drainage basins. Changes in these ratios
in size-fractionated samples, assuming a constant and relatively homogeneous source in the tills, reflect the relative
chemical stabilities of the clastic components: calcite and
feldspars are more soluble relative to dolomite and quartz,
and are therefore less likely to survive in a near-surface
weathering environment.

Stable isotopes
Investigation of the stratigraphic variation in stable oxygen
and carbon isotopes of lacustrine sediment has led to important paleohydrological observations and paleoclimatic interpretations in many modern and ancient lakes (see reviews in
Pearson and Coplen, 1978; Buchardt and Fritz, 1980; Kelts
and Talbot, 1989; Swart et al., 1993). Stable isotope analyses
have also been instrumental in helping understand the complex genesis and diagenesis of endogenic and authigenic minerals in lake sediments (e.g. Talbot and Kelts, 1986; Rosen et
al., 1989; Perkins et al., 1994). Historically, inorganic carbonates (mainly calcite and aragonite) and biological
endogenic carbonates (ostracode shells) have been the main
targets for paleoenvironmental research. More recently, oxygen isotopes of noncarbonate organic components have been
shown to be valuable in supplementing the paleoenvironmental information that can be deduced (Edwards and Fritz, 1988;
Edwards and McAndrews, 1989; Edwards, 1993; Wolfe
et al., 1996; Wolfe and Edwards, 1997). These combined
approaches, together with trace element measurements, offer
a powerful means of deducing past water levels, salinities,
temperatures, and organic productivity in lake basins.
Although the general principles of oxygen and carbon isotopes are reasonably well known and straightforward, interpretation of data from nonmarine settings can be complex.
The oxygen isotopic composition of endogenic minerals
depends on the isotopic composition and temperature of the
solution at the time of precipitation. The isotopic composition
of the lake water, in turn, depends mainly on the composition
of inflowing waters and the residence time of the water in the
lake. In lakes characterized by open hydrological systems,
residence times are generally low, the waters have not been
subjected to isotopic evolution within the basin, and the isotopic composition differs little from that of inflowing waters.
However, in closed hydrological settings, residence time
increases and lake waters can undergo significant enrichment
of 1 8 0 , because the lighter 160isotope is preferentially incorporated into the gaseous phase during evaporation. Because
fractionation of 1 8 0 is temperature dependent between the
mineral phase and aqueous phases, but independent of temperature between the aqueous phase and endogenic organic
matter, it is possible to infer lake water paleotemperatures by
examining the ratios from both mineral and organic matter.
The factors controlling the ratio of 13cto 12cin the -precipitated carbonate are not directly related to hydrology and
evaporation/precipitation ratios in the lacustrine basin. However, the preferential uptake of 12cby aquatic vegetation and
preferential outgassing of 12c-rich COPfrom the lake surface
during long residence times lead to a marked covariance of
6I3c and 6180 in closed basins.
-

Despite these demonstrated applications to


paleolimnology, there have been surprisingly few isotopic
studies on lake sediments in the Canadian Prairies. Last et al.
(1994) used the 613c and 6180 of endogenic inorganic

GSC Bulletin 534

magnesian calcite and ostracode shells to track the hydrological fluctuations of the 12 000-year long record in the Lake
Manitoba basin. Results show a complex hydrochemical and
hydrological history due to multiple and shifting water
sources, regional climate changes, and repeated alternations
between open and closed basin conditions. Boden (1985) and
Last and Slezak (1986) summarized the isotopic record of
endogenic aragonite in Deadmoose Lake. They found relatively minor hydrological variations but identified several
periods of increased primary organic productivity during the
past 2500 years in this saline, meromictic lake. Van
Stempvoort et al. (1993) were similarly able to identify episodes of enhanced aridity and increased productivity andlor
warmth in the 2500-year long aragonite-laminated record
from Redberry Lake, another deep-water saline lake northwest of Saskatoon. Most recently, Padden (1996) examined
the stableisotopes of both carbonate and cellulose in the sediments of Chappice, Oro, and Killarney lakes. The complex
endogenic carbonate mineralogies of Chappice and Oro
lakes, and the occurrence of relatively large proportions of
allogenic carbonate in Killarney Lake, hampered her use of
quantitative paleothermometry in these lakes. Nevertheless,
she recognized periods of hydrological aridity that were generally consistent with other lithostratigraphic parameters
investigated in these basins.

GREAT PLAINS GEOLIMNOLOGY:


THE DECADE AHEAD
Accordingto the proverb, the best things are the most diff~cult.
-Plutarch, AD 46-120
Geolimnology is a diverse subdiscipline incorporating
many of the same basic precepts as lirnnology but with an
emphasis on the genesis and diagenesis of sediments in the
basin and on the stratigraphy of the basin fill. It is clear that
enormous advances have been made this century in both characterizing and understanding lakes in the northern Great
Plains. The past several decades have witnessed explosive
growth in scientific interest and research on these lakes due,
in part, to significant technological advances involving sarnple and core acquisition, chronostratigraphy, and analytical
techniques. However, much of today's ongoing geolimnological and paleolirnnological research is a direct outgrowth
of principles and concepts established in the early part of the
twentieth century. Indeed, the first 'academic' (as opposed to
applied) investigation of the sediments in these lakes did not
appear until nearly 50 years after the start of mining and commercial exploitation of the deposits. This initial emphasis, by
mainly economic geoscientists on the preserved salts and on
the most concentrated lakes within the vast range of salinity
types of the region, established a pattern that largely
continues today. We know considerably more about the physical and chemical processes, the modern sediments, and the
Holocene stratigraphies in the hypersaline lakes and playas
than we do about those lakes with more moderate salinities.
Future geolimnological research must be directed toward the
moderate- and low-salinity lakes of the region to fill this
information void.

Foremost among the accomplishments of the past century


is an appreciation of the great breadth of water compositions
and concentrations exhibited by these lakes. From this broad
spectrum of water-chemistry types, we now realize that the
inorganic sediment records of the basins are characterized not
by the relatively simple mineral assemblages often present in
many other lakes of the world, but rather by a diverse and
sometimes exceedingly complex assortment of endogenic,
authigenic, and allogenic minerals. This striking diversity,
both spatial and temporal, is of such fundamental importance
that it pervades virtually every other geolimnological consideration. The presence of this wealth of endogenic and
authigenic inorganic constituents, coupled with our understanding of the thermodynamic conditions of genesis, provide us the opportunity to evaluate, often in explicit
quantitative detail, past water compositions and ionic concentrations. This is of particular importance because many of
the biological proxy indicators commonly used by
paleolirnnologists (e.g. siliceous microfossils, ostracodes,
chironomids, pollen) are ineffective at elevated salinities, are
complacent in the prairie environment, or experience significant postdepositional alteration in the brines (Evans, 1993).
The allogenic fraction of the sediment permits investigators
to make important, albeit qualitative, observations about the
weatheringconditions (and hence the climatic regime) in the
catchment, water-level fluctuations in the basin, and the
nature of the erosional processes and transportation agents in
the lake and drainage basin.

A vital consideration in approaching stratigraphic records


and attempting to interpret past climatic, hydrological, and
environmental conditions in the Great Plains is the establishment of quantitative information about modern sedimentary
processes. The axiom 'the present is the key to the past' is of
particular relevance when the processes of saline karst, mud
diapirism, and sedimentation rates are considered for these
lakes. Quantitative investigations of each of these processes
have provided important insight into conditions responsible
for ancient evaporite and lacustrine deposits. Similarly, documentation of the reaction mechanisms for dolomite formation
in these lakes has considerably advanced our understanding
of nonmarine genesis of this important mineral and has contributed to helping resolve the 'dolomite problem'.
What remains for the lacustrine sedimentologist and geochemist to do in western Canada? Although our understanding of the broad regional trends in water composition is
acceptable, minimal progress has been made on identifying
and quantifying the source of dissolved components for the
basins. Indeed, it is disconcerting that, even in several of the
best studied lakes (e.g. Ingebrigt, Kenosee, Oro, Quill,
Waldsea, Winnipeg), derivation of the solutes is largely
unknown. A broader attack on this problem is required,
including more work along the lines of Wallick (1981),
Hendry et al. (1986), and Remenda et al. (1994).
Coupled with this geochemical explication of the lake
waters, it is essential that a more complete understanding of
the hydrological budgets in these lakes be achieved. On a
broad, regional basis, the hydrogeology of the northern Great
Plains is well advanced (Remenda and Birks, 1999). Nonetheless, detailed site-specific investigations of lacustrine

W.M. Last

systems are few, and important aspects of groundwater-lake


system interaction have received little attention. Our ability
to interpret both inorganic sediments and organic remains in
many of these basins depends largely upon a suitable understanding of the hydrological system (Lemmen et al., 1997). In
particular, we must have improved awareness of how the
hydrology of these basins responds to changes in climate. It is
clear from the few detailed investigations that have been carried out (Freeze, 1969; Hammer, 1990; Padden, 1996; Last
and Teller, 1997; Birks and Remenda, 1999) that simple
hydrological perceptions like 'wet climate-high lake
level-freshwater; dry climate-low lake level-saline water'
do not apply to many basins in the Great Plains. Indeed, surprisingly little effort has been made to quantitatively document or calibrate the sedimentological and geochemical
response of individual lacustrine basins and their catchments
to climate-induced (or anthropogenic) changes in vegetation
cover, morphology, runoff, and nutrient availability. Davis
(1990, p. 18) commented that "Many intriguing hypotheses
can be formulated around the general question: do lake
responses track changes in the catchment?'In the Great
Plains region, these hypotheses remain largely untested.
There are several underexploited geolirnnological techniques and areas of investigation which should be scrutinized
in future projects. As discussed above, the use of stable isotopes to better understand the hydrology, composition, and
productivity of the lakes holds great promise. Mineral magnetic data, considered by many to be basic and essential information for any lake sediment study (e.g. Thompson and
Oldfield, 1986; Colman, 1996) have been collected on only
two lacustrine sequences in the entire northern Great Plains
(Lake Manitoba and Lake Winnipeg). These data have potential to contribute much to our understanding of sediment
sources, erosion patterns, land-use changes, and water-level
fluctuations. Similarly, charcoal analysis, commonly used
elsewhere to decipher natural and anthropogenic changes in
drainage basin vegetation and soils, has been overlooked in
most studies of lakes in the Great Plains (Campbell, 1997,
1998). Although interpretive models are complex, the use of
pore water to gauge past limnological conditions has been
only rarely attempted in this region (Teller and Last, 1979;
Buhay and Betcher, 1998). This technique holds considerable
promise for supplementing paleochemical information in
basins that do not have suitable endogenic mineral precipitates. The relatively new techniques of geochemical and isotopic analyses of fluid inclusions in salts have been used
successfully in other saline lake sequences to interpret
paleotemperature, humidity, and brine composition (Roberts
and Spencer, 1995;Roberts et al., 1997) but have not yet been
applied to the evaporites of the Great Plains.
In general, the greatest progress in understanding past
changes within lacustrine environments, and interpreting
these changes with respect to either short- or long-term
extrinsic factors such as climate, basin evolution, or anthropogenic alteration, comes from integrative and co-ordinated
multidisciplinary efforts (Chivas and Bowler, 1986;Oldfield,
1996). This approach has made significant contributions to
our understanding of the northern Great Plains within the past
decade (Lemrnen et al., 1997; Vance et al., 1998; Todd et al.,

1998), although, by necessity, progress has been slow. To a


great extent, the challenge remaining for geolimnologists and
paleolimnologists in the northern Great Plains is to continue
the application of this multidisciplinary strategy on a greater
diversity of basins within this large region of North America.

ACKNOWLEDGMENTS
In comparing various authors with one another, I have discovered that some of the gravest and latest writers have tran-

scribed, word for word, from former works, without making


acknowledgment.
-Pliny the Elder, Natural History, Book I.
Dedication, Section 22
Much of this manuscript was prepared for, and presented
as parts of, guidebooks for two field trips run in conjunction
with a conference on 'Sedimentary and Paleolirnnological
Records of Saline Lakes', held in Saskatoon in 1991 (Last,
1991a, b). I thank the organizers of this conference for permission to use this material. I am also indebted to the students
and co-researchers with whom I have worked on these lakes
over the years and who have provided much of the data and
observations summarized here. Special thanks go to D. Bell,
S. Boden, A. Carter, G. Carlese, M. Egan, M.E. Hines,
L. Kovac, R.M. Lent, E. Lockhart, W.B. Lyons, A. Ross,
L. Sack, G. Sayers, B. Schiitt, T. Schweyen, Y. Shang,
L. Slezak, J. Wei, and A. Whitney.
Finally, since the beginning of the Palliser Project in
1991, the 'lakes' portion of the research effort has been, in
two words, Bob Vance. Bob provided the leadership, direction, and co-ordination, as well as scientific and technical
expertise, essential to the success of the project. I do not know
what the future holds for lake sediment research in western
Canada; we shall see in coming years. But undertaking
paleoenvironmental work in the region without Bob is difficult to imagine; his contributions will be sorely missed.

REFERENCES
Aber, J.S.
1993: Geomorphology and structural genesis of the Dirt Hills and Cactus
Hills, southern Saskatchewan; in Glaciotectonics and Mapping
Glacial Deposits (Proceedings of the INQUA Commission on
Formation and Properties of Glacial Deposits), (ed.) J.S. Aber;
Canadian Plains Research Centre, University of Regina, Regina,
Saskatchewan, Canadian Plains Proceedings, no. 25-1, p. 9-35.
Adams, G.D.
1988: Wetlands of the Prairies of Canada; in Wetlands of Canada;
National Wetlands Working Group, Canada Committee on Ecological Land Classification, Ecological Land Classification Series,
No. 24, Environment Canada, Ottawa, Ontario, p. 155-198.
Adams, G.D., Didiuk, A.B., Pochailo, D., and Eskowich, K.
1992: Impact of drought and land use on prairie potholes in the Antler
Municipality, Saskatchewan; irz Aquatic Ecosystems in Semi-Arid
Regions, (ed.) R.D. Robarts and M.L. Bothwell; National
Hydrology Research Institute Symposium No. 7, Environment
Canada, Saskatoon, Saskatchewan, p. 27-40.
Agrawal, Y.C., McCave, I.N., and Riley, J.B.
1991: Laser diffraction size analysis; in Principles, Methods, and Application of Particle Size Analysis, (ed.) J.P. Syvitski; Cambridge
University Press, New York, p. 119-128.

GSC Bulletin 534


Anadbn, P., Cabrera L.I., and Kelts, K. (ed.)
1991: Lacustrine facies analysis; Blackwell Scientific Publications,
Oxford, United Kingdom, 317 p.
Arndt, J.L. and Richardson, J.L.
1992: Carbonateand gypsum chemistry in saturated, neutral pH soil environments; in Aquatic Ecosystems in Semi-Arid Regions, (ed.)
R.D. Robarts and M.L. Bothwell; National Hydrology Research
Institute Symposium No. 7, Environment Canada, Saskatoon,
Saskatchewan,p. 179-187.
Barica, J.
1975: Geochemistry and nutrient regime of saline eutrophic lakes in the
Erickson-Elphinstone district of southwesternManitoba; Canadian
Fisheries and Marine Services, Research Development Technical
Report511,82 p.
Barry, G.S.
1986: Sodium sulphate; Canadian Mining Journal, v. 107, p. 77.
Barry, G.S., Boucher, J.A., and Prud'homme, M.
1985: Other industrial minerals; Canadian Mining Journal, v. 106,
p. 82-83.
Batt, B.D., Anderson, M.G., Anderson, C.D., and Caswell, F.D.
1989: The use of prairie potholes by North American ducks; in Northern
Prairie Wetlands, (ed.) A.G. van der Valk; Iowa State Press, Ames,
Iowa, p. 204-227.
Beaty, C.B.
1971: Climate change: some doubts; Geological Society of America
Bulletin, v. 82, p. 1395-1397.
Beaudoin, A.B.
1993: A compendium and evaluation of postglacial pollen records in
Alberta; Canadian Journal of Archeology, v. 17, p. 92-1 12.
Bell, R.
1874: Report on the country between Red River and the South
Saskatchewan; Geological Survey of Canada, Report of Progress,
1873-1874, p. 66-91.
Betclier, R., Grove, G., and Pupp, C.
1995: Groundwater in Manitoba: hydrogeology, quality concerns, management; Environment Canada, National Hydrology Research
Institute, Contribution No. CS-93017, 47 p.
Bierhuizen, J.F.H. and Prepas, E.E.
1985: Relationships between nutrients, dominant ions, and phytoplankton
standing crop in prairie saline lakes; Canadian Journal of Fisheries
and Aquatic Sciences, v. 42, p. 1588-1594.
Binyon, E.
1952: North Dakota sodium sulfate deposits; United States Bureau of
Mines, Report of Investigation No. 4880,41 p.
Bird, J.B.
1980: The Natural Landscapes of Canada, a Study in Regional Earth
Science; John Wiley & Sons Canada Limited, Toronto, Ontario,
260 p. (second edition).
Birks, S.J. and Remenda, V.H.
1999: Hydrogeological investigation of Chappice Lake, southeastern
Alberta: groundwater inputs to a saline basin; Journal of
Paleolimnology, v. 21, p. 235-255.
Bluemle, J.P.
1991: The face of North Dakota; North Dakota Geological Survey,
Education Series 21, 177 p.
1992: Vebl~ns,hydrodynamic blowouts, and other unusual geologic features in North Dakota; North Dakota Geological Survey, Newsletter, v. 19, p. 7-.11.
1993: Hydrodynamic blowouts in North Dakota; in Glaciotectonics and
Mapping Glacial Deposits (Proceedings of the MQUA Comrnission on Formation and Properties of Glacial Deposits, (ed.)
J.S. Aber; Canadian Plains Research Centre, University of Regina,
Regina, Saskatchewan, Canadian Plains Proceedings, v. 25-1,
p. 259-266.
Boden, S.E.
1985: A chemical and sedi~nentologicalstudy of Deadmoose Lake,
Saskatchewan; B.Sc. thesis, University of Manitoba, Winnipeg,
Manitoba, 79 p.
Boulton, G.S. and Caban, P.
1995: Goundwater flow beneath ice sheets: part I1 - its impact on
glaciotectonic structures and moraine formation; Quaternary
Science Reviews, v. 14, p. 563-587.
Braitsch, 0.
1971: Salt Deposits, Their Origin and Composition; Springer-Verlag,
New York, 297 p.

Broughton, P.L.
1984: Sodium sulphate deposits of western Canada; in The Geology of
Industrial Minerals in Canada, (ed.) G.R. Guillet and W. Martin;
Canadian Institute of Mining and Metallurgy, Special Volume 29,
p. 195--200.
Brown, I.c.
1967: Groundwater in Canada; Geological Survey of Canada, Economic
Geology Report 24,228 p.
Brunskill, G.J. and Graham, B.W.
1979: The offshore sediments of Lake Winnipeg; Canadian Fisheries and
Marine Services, Manuscript Report No. 1540.75 p.
Buchardt, B. and Fritz, P.
1980: Environmental isotopes as environmental and climatological indicators; in Handbook of Environmental Isoto~eGeochemistrv.
volume I, The Terrestrial Environment, (ed.) P. Fritz and J . c ~ :
Fontes; Elsevier, New York, p. 473-504.
Buhay, W.M. and Betcher, R.N.
1998: Paleohydrologicimplications of ' 8 0 enriched Lake Agassiz water;
Journal of Paleolimnology,v. 19, p. 285-296.
Callender. E.
1968: ~ h e ~ o s t ~ l a csedimentology
ial
of Devils Lake, North Dakota, Ph.D.
thesis, University of North Dakota, Grand Forks, North Dakota, 3 12
P.
Cameron, D.R.
1986: Saline waters: agricultural uses; in Evaluating Saline Waters in a
Plains Environment, (ed.) D.T. Waite; Canadian Plains Research
Centre, University of Regina, Regina, Saskatchewan, Canadian
Plains Proceedings, v. 17, p. 75-100.
Campbell, C.
1997: Postglacial geomorphic response and environmental change in
southeastern Alberta, Canada; Ph.D. thesis, University of Alberta,
Edmonton, Alberta, 187 p.
1998: Postglacial evolution of a fine-grained alluvial fan in the northern
Great Plains, Canada; Palaeogeography, Palaeoclimatology,
Palaeoecology, v. 139, p. 233-249.
Campbell, C., Can~pbell,I.D., and Hogg, E.H.
1994: Lake area variability across a climatic and vegetational transect in
southeasternAlberta, Canada; Geographiephysiqueet Qnaternaire,
V. 48, p. 207-212.
Campbell, I.D., Campbell, C., Apps, M.J., Rutter, N.W.,
and Bush, A.B.G.
1998: Late Holocene -1500 yr climatic periodicities and their implications; Geology, v. 26, p. 471473.
Canadian National Comnlittee for the International Hydrologic
Decade
1978: Hydrologic atlas of Canada; Fisheries and Environment Canada,
Ottawa, 75 p.
Carpenter, A.B.
1978: Origin and chemical evolution of brines in sedimentary basins;
Oklahoma Geological Survey, Circular 79, p. 60-77.
Chilingar, G.V., Zenger, D.H., Bissell, H.J., and Wolf, K.H.
1979: Dolomites and dolomitization; in Diagenesis in Sediments and
Sedimentary Rocks, (ed.) G. Larsen and G.V. Chilingar; Developments in Sedimentology 25A, Elsevier, Amsterdam, Netherlands,
p. 425-526.
Chivas, A.R. and Bowler, J.M.
1986: Introduction - The SLEADS Project; Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 54, p. 3-6.
Christiansen, E.A.
1967a: Collapse structures near Saskatoon, Saskatchewan; Canadian Journal of Earth Sciences, v. 4, p. 757-767.
1967b: Preglacial valleys in southern Saskatchewan; Saskatchewan
Research Council, Saskatoon, Saskatchewan, Map 3, scale
1520 640.
1971: Geology of the Crater Lake collapse structure in southeastern
Saskatchewan; Canadian Journal of Earth Sciences, v. 8,
p. 1505-1513.
1973: Quaternary geology and its application to engineering practice in
the Saskatoon-Regina-Watrous area, Saskatchewan; Geological
Association of Canada, Guidebook, Field Trip C, 81 p.
1979: The Wisconsinan deglaciation of southern Saskatchewan and adjacent areas; Canadian Journal of Earth Sciences, v. 16, p. 913-938.
Christiansen, E.A. and Whitaker, S.H.
1976: Glacial thrusting of drift and bedrock; in Glacial Till: An Interdisciplinary Study, (ed.) R.F. Legget; The Royal Society of Canada,
Special Publication No. 12, p. 121-130.

W.M. Last
Christiansen, E.A., Gendzwill, D.J., and Meneley, W.A.
1982: Howe Lake: a hydrodynamic blowout structure; Canadian Journal
of Earth Sciences, v. 19, p. 1122-1 139.
Clayton, L. and Cherry, J.A.
1967: Pleistocenesuperglacial and ice-walled lakes of west-central North
America; in Glacial Geology of The Missouri Coteau and Adjacent
Areas, (ed.) L. Clayton and T.F. Freers; North Dakota Geological
Survey, Miscellaneous Series 30, p. 47-52.
Clayton, L. and Freers, T.F.
1967: Road Log; in Glacial Geology of TheMissouri Coteau and Adjacent
Areas, (ed.) L. Clayton and T.F. Freers; North Dakota Geological
Survey, Miscellaneous Series 30, p. 1-19.
Cole, G.A.
1975: Textbook of Limnology; Mosby, St. Louis, Missouri, 283 p.
Cole, L.H.
1926: Sodium sulfate of western Canada - occurrence, uses and technology; Canadian Department of Mines, Publication No. 646, 155 p.
1930: The salt industry of Canada; Canadian Department of Mines, Publication No. 716, 116 p.
Colman, S.M. (ed.)
1996: Continental drilling for paleoclimatic records: recommendations
from an international workshop, Potsdam, June 30-July 2, 1995;
PAGES (Past Global Changes), Bern, Switzerland, Workshop
Report, Series 96-4, 104 p.
Davis, R.B.
1990: The scope of Quaternary paleolimmnology; in Paleolimnology and
the Reconstruction of Ancient Environments, (ed.) R.B. Davis;
Kluwer Academic Publishers, Dordrecht, Netherlands, p. 1-24.
Dean, W.E.
1981: Carbonate minerals and organic matter in sediments of modern
north temperate hard-water lakes; in Recent and Ancient
Nonmarine Depositional Environments: Models for Exploration,
(ed.) F.G. Ethridge and R.M. Flores; Society of Economic Palcontologists and Mineralogists, Special Publication No. 31,
p. 213-231.
Dean, W.E. and Fouch, T.D.
1983: Lacustrine; Chapter 2 in Carbonate Depositional Environments,
(ed.) P.A. Scholle, D.G. Bebout and C.H. Moore; American
Association of Petroleum Geologists, Memoir 33, p. 98-130.
Dean, W.E. and Gorham, E.
1976: Major chemical and mineral components of profundal surface sediments in Minnesota Lakes; Limnology and Oceanography, v. 21,
p. 259-284.
Dean, W.E. and Megard, R.O.
1993: Environment of deposition of CaC03 in Elk Lake, Minnesota; in
Elk Lake Minnesota: Evidence for Rapid Climate Change in the
North-Central United States, (ed.) J.P. Bradbury and W.E. Dean;
Geological Society of America, Special Paper 276, p. 97-1 13.
Dean, W.E. and Stuiver, M.
1993: Stable carbon and oxygen isotope studies of the sediments of Elk
Lake, Minnesota; in Elk Lake Minnesota: Evidence for Rapid
Climate Change in the North-Central United States, (ed.)
J.P. Bradbury and W.E. Dean; Geological Society of America, Special Paper 276, p. 163-180.
De Deckker, P. and Last, W.M.
1989: Modern, non-marine dolomite in evaporitic playas of western
Victoria, Australia; Sedimentary Geology, v. 64, p. 223-238.
Desai, K. and Moore, E.J.
1969: Equivalent NaCl determination from ionic concentrations; Log
Analyst, v. 10, p. 12-21.
Doering, E.J.
1982: Water and salt movement in prairie soils; in Proceedings of Soil
Salinity Conference, Lethbridge, Alberta, (ed.) B.L. Colgan;
Alberta Agriculture, Edmonton, Alberta, p. 55-78.
Donovan, J.J.
1992: Geochemical and hydrological dynamics of evaporativegroundwater-dominated lakes of Montana and North Dakota; Ph.D. thesis,
Pennsylvania State University, University Park, Pennsylvania,
251 p.
1994: On the measurement of reactive mass fluxes in evaporative groundwater-sourcelakes; in Sedimentology and Geochemistry of Modern
and Ancient Saline Lakes, (ed.) R. Renaut and W.M. Last; Society
of Economic Paleontologists and Mineralogists, Special
Publication No. 50, p. 33-50.

Donovan, J.J. and Rose, A.W.


1994: Geochemical evolution of lacustrine brines from variable-scale
groundwater circulation; Journal of Hydrology, v. 154, p. 35-62.
Downey, J.S.
1969: -Geology of northeastern North Dakota; North Dakota Geological
Survey. Miscellaneous Series 39.30 D.
Dreimanis, and Vagners, U.J.
1971: Bimodal distribution of rock and mineral fragments in basal till; in
Till - A Symposium, (ed.) R.P. Goldthwait; Ohio State University
Press, Columbus, Ohio, p. 237-250.
Drever, J.I.
1988: The Geochemistry of Natural Waters; Prentice Hall, Englewood
Cliffs, New Jersey, 437 p. (second edition).
Driver, E.A.
1965: Limnological aspects of some saline lakes in west-central
Manitoba; Journal of Fisheries Research Board of Canada, v. 22,
p. 1165-1173.
Driver, E.A. and Pedden, D.G.
1977: The chemistry of surface water in the prairie ponds; Hydrobiologia,
v. 53, p. 3 3 4 8 .
Dyke, A.S. and Prest, V.K.
1987: Late Wisconsinan and Holocene retreat of the Laurentide Ice Sheet;
Geological Survey of Canada, Map 1702A, scale 1:5 000 000.
Eatock, W.H.
1987: The Big Quill Lake sulphate of potash project; in Economic
Minerals of Saskatchewan, (ed.) C.F. Gilboy and L.W. Vigrass;
Saskatchewan Geological Society, Regina, Saskatchewan, Special
Publication No. 8, p. 206-210.
Edwards, T.W.D.
1993: Interpretingpast climate from stable isotopes in continental organic
matter; in Climatic Change in Continental Isotopic Records, (4.)
P.K. Swart, K.C. Lohmann, J. McKenzie, and S. Savin; American
Geophysical Union, Geophysical Monograph 78, Washington,
D.C., p. 333-341.
Edwards, T.W.D. and Fritz, P.
1988: Stable isotope paleoclimate records for southern Ontario, Canada:
comparison of results from marl and wood; Canadian Journal of
Earth Sciences, v. 25, p. 1397-1406.
Edwards, T.W.D. and McAndrews, J.H.
1989: Paleohydrologyof a Canadian Shield lake inferred from
in sediment cellulose; Canadian Journal of Earth Sciences, v. 26,
p. 1397-1406.
Egan, D.M.
1984: Mineralogical aspects of hypersaline lakes in southern
Saskatchewan; M.Sc. thesis, University of Manitoba, Winnipeg,
Manitoba, 192 p.
Ehrlich, W.A. and Rice, W.M.
1955: Post-glacial weathering of Mankato till in Manitoba; Journal of
Geology, v. 63, p. 527-537.
Einsele, G.
1992: Sedimentary Basins; Springer-Verlag,Berlin, Germany, 483 p.
Environment Canada
1975: Water quality data - Alberta, 1961-1973; Environment Canada,
Inland Waters Directorate, Water Quality Branch, 650 p.
1980: Detailed surface water quality data - Manitoba, 1974-1977;
Environment Canada, Inland Waters Directorate, Water Quality
Branch, 356 p.
1984: Detailed surface water quality data - Saskatchewan, 1980-1981;
Environment Canada, Inland Waters Directorate, Water Quality
Branch, 156 p.
Eugster, H.P. and Hardie L.A.
1978: Saline lakes; in Lakes - Chemistry, Geology, Physics, (ed.)
A. Lerman; Springer Verlag, New York, p. 237-293.
Eugster, H.P. and Kelts, K.
1983: Lacustrine chemical sediments; in Chemical Sediments and
Geomorphology, (ed.) A.S. Goudie and K. Pye; Academic Press,
London, United kngdom, p. 321-368.
Evans, D.J.A., Lemmen, D.S., and Rea, B.R.
in press: Glacial land systems of the southwest Laurentide Ice Sheet: modern
Icelandic analogues; Journal of Quaternary Science.
Evans, J.C. and Prepas, E.E.
1996: Potential effects of climate change on ion chemistry and
phytoplankton communities in prairie saline lakes; Limnology and
Oceanography, v. 41, p. 1063-1076.

GSC Bulletin 534


Evans, M.S.
1993: Paleolimnological studies of saline lakes; Journal of
Paleolimnology, v. 8, p. 97-101.
Freeze, R.A.
1969: Regional groundwater flow - Old Wives Lake drainage basin,
Saskatchewan; Canadian Inland Waters Directorate, Scientific
Series No. 5,234 p.
Fritz, S.C.
1996: Paleolilnnological records of climatic change in North America;
Limnology and Oceanography,v. 41, p. 882-889.
Gendzwill, D.J. and Hajnal, Z.
1971: Seismic investigation of the Crater Lake collapse structure in southeastern Saskatchewan; Canadian Jou~nalof Earth Sciences, v. 8,
p. 1514-1524.
Ghebre-Egziabhier, K. and St. Arnaud, R.J.
1983a: Carbonate mineralo~yof lake sediments and surrounding, soils I.
Blackstrap Lake; Canadian Journal of Soil science, v. 63,
p. 245-257.
1983b: carbonate mineralogy of lake sediments and surrounding soils 2.
the Ou'Ap~ellelakes. Canadian Journal of Soil Science. v. 63.
p. 259-26'9.
Gierlowski-Kordesch. E. and Kelts. K. (ed.)
e
Volume 1; Cambridge
1994: Global ~eologicalRecord of ~ a k Basins:
University Press, New York, 427 p.
Gollop, J.B.
1963: Wetland inventoriesin western Canada;Transactions,6th Congress,
International Union of Game Biologists, Bournemoutl~,United
Kingdom, p. 249-264.
Gorham, E., Dean, W.E., and Sanger, J.E.
1983: The chemical composition of lakes in the north-central United
States; Limnology and Oceanography,v. 28, p. 287-301.
Goudie, A.S. and Cook, R.U.
1984: Salt efflorescences and saline lakes, a distributional analysis;
Geoforum, v. 15, p. 563-582.
Govett, G.J.S.
1958: Sodium sulfate deposits in Alberta; Alberta Research Council,
Preliminary Report 58-5.34 p.
Gravenor, C.P. and Kupsch, W.O.
1959: Ice-disintegrationfeatures in western Canada; Journal of Geology,
V. 67, p. 48-67.
Groom, H.D.
1985: Surficial geology and aggregate resources of the Fisher Branch
area; Manitoba Energy and Mines, AggregateReport AR84-2,33 p.
Grossman, LG.
1949: The sodium sulfate deposits of western North Dakota; North Dakota
Geological Survey, Report of Investigations No. 1,66 p.
1968: Origin of sodium sulfate deposits of the northern Great Plains of
Canada and the United States; United States Geological Survey,
Professional Paper 600-B, p. 104-109.
Gustavson, T.C. and Holliday, V.T.
1992: Playa basin development, southern High Plains, Texas and New
Mexico; Geological Society of America, Annual Meeting,
Abstracts with Program, v. 24, p. 122.
Hlkanson, L.
1977: The influence of wind, fetch, and water depth on the distribution of
sediments in Lake Vanern, Sweden; Canadian Journal of Earth
Sciences, v. 14, p. 397-412.
HBkanson, L. and Jansson, M.
1983: Principlesof Lake Sedimentology;Springer-Verlag,Berlin, 316 p.
Hammer, U.T.
1978a: The saline lakes of Saskatchewan,I. background and rationale for
saline lakes research; Internationale Revue der Gesamte
Hydrobiologie, v. 63, p.173-177.
1978b: The saline lakes of Saskatchewan, III. chemical characterization;
Internationale Revue der Gesamte Hydrobiologie, v. 63,
p. 311-335.
1981: Primary production in saline lakes - a review; Hydrobiologia,v. 66,
p. 701-743.
1984: The saline lakes of Canada; in Lakes and Reservoirs, (ed.)
F.B. Taub; Ecosystems of the World 23, Elsevier, Amsterdam,
Netherlands, p. 510-540.
1986: Saline Lake Ecosystems of the World; Dr. W. Junk Publishing,
Dordrecht, Netherlands, 616 p.

Hammer, U.T. (cont.)


1990: The effects of climate change on the salinity, water levels and biota
of Canadian prairie saline lakes; Verhandlungender Internationalen
Vereinigung fur Theoretische und Angewandte Limnologie, v. 24,
p. 321-326.
Hammer, U.T. and Haynes, R.C.
1978: The saline lakes of Saskatchewan, 11. locale, hydrology, and other
physical aspects; Internationale Revue der ~ e s a r n t~ydrobiolo~ie,
e
v. 63, p. 179-203.
Hammer, U.T: and Parker, J.
1984: Limnology of a perturbed highly saline Canadian lake; Archiv fur
Hydrobiologie, v. 102, p. 31-42.
Hammer, U.T., Shamess, J., and Haynes, R.C.
1983: The distribution and abundance of algae in saline lakes of
Saskatchewan, Canada; Hydrobiologia, v. 105, p. 1-26.
Handford, C.R.
1981: A process-sedimentary framework for characterizing recent and
ancient sabkhas; Sedimentary Geology, v. 30, p. 255-265.
Hardie, L.A. and Eugster, H.P.
1971: The evolution of closed basin brines; Mineralogical Society of
America, Special Publication 3, p. 273-290.
Hardie, L.A., Smoot, J.P., and Eugster, H.P.
1978: Saline lakes and their deposits: a sedimentological approach; in
Modern and Ancient Lake Sediments, (ed.) A . Matter and
M.E. Tucker; International Association of Sedimentologists,
Special Publication No. 2, p. 7-42.
Hartland-Rowe, R.
1966: The fauna and ecology of temporary pools in western Canada;
Verhandlungen der Internationalen Vereinigung fur Theoretische
und Angewandte Limnologie, v. 16, p. 577-584.
Haskell, B.J., Fritz, S.C., and Engstrom, D.R.
1996: Late Quaternary paleohydrology in the North American Great
Plains inferred from the geochemistry of endogenic carbonate and
fossil ostracodes from Devils Lake, North Dakota; Palaeogeography, Palaeclimatology, Palaeoecology, v. 124, p. 179-193.
Haynes, R.C. and Hammer, U.T.
1978: The saline lakes of Saskatchewan, IV. primary production of
phytoplankton in selected saline ecosystems; Internationale Revue
der Gesamte Hydrobiologie, v. 63, p. 337-351.
Hem, J. D.
1985: Study and interpretation of the chemical characteristics of natural
waters, United States Geological Survey, Water-Supply Paper
2254,263 p. (third edition).
Henderson, P. and Last, W.M.
1998: Holocene sedimentation in Lake Winnipeg, Manitoba, Canada:
implications of compositional and textu~alvariations; Journal of
Paleolimnology,v. 19, p. 265-284.
Hendry, M.J. and Schwartz, I?.
1982: Hydrogeology of saline seeps; in Proceedings of Soil Salinity
Conference, Lethbridge, Alberta, (ed.) B.L. Colgan; Alberta Agriculture, Edmonton, Alberta, p. 25-53.
Hendry, M.J., Cherry, J.A., and Wallick, E.I.
1986: Origin and distribution of sulphate in a fractured till in southern
Alberta, Canada; Water Resources Research, v. 22, p. 45-61.
Herrington, R., Johnson, B., and Hunter F. (ed.)
1997: Responding to global climate change in the Prairies; Volume 111 of
the Canada Country Study: Climate Impacts and Adaptation,
Environment Canada, Prairie and Northern Region, 75 p.
Hind, H.Y.
1861: Narrative of the Canadian Red River Exploring Expedition of 1857
and of the Assinniboine and SaskatchewanExploring Expedition of
1858; MacLear, Toronto, Ontario, 12 p.
Holser, W.T.
1979: Mineralogy of evaporites; in Marine Minerals, (ed.) R.G. Burns;
Mineralogical Society of America, Short Course Notes, v. 6,
p. 21 1-294.
James F. MacLaren Limited
1980: Mineral aggregate study of the southern Interlake region; Manitoba
Department of Energy andMines, Open File Report OF80-2,51 p.
Jones, B.F. and Bowser, C.J.
1978: The mineralogy and related chemistry of lake sediments; in Lakes Chemistry, Geology, Physics, (ed.) A. Lerman; Springer Verlag,
New York, p. 179-235.

W.M. Last

Jones, B.F. and van Denburgh, A.S.


1966: Geochemical influences on the chemical character of closed lakes;
in Symposium of Garda, Hydrology of Lakes and Reservoirs; International Association for Scientific Hydrology, Publication 70,
p. 438-446.
Kehew, A.E. and Lord, M.L.
1989: Glacial-lake spillways of the central interior plains, CanadaU.S.A.; Canadian Geographer, Canadian Landform Examples - 12,
v. 33, p. 274-277.
Kehew, A. E. and Teller, J. T.
1994: History of late glacial runoff along the southwestern margin of the
Laurentide ice sheet; Quaternary Science Reviews, v. 13,
p. 859-877.
Keller, L.P., McCarthy, G.J., and Richardson, J.L.
1986a: Laboratory modeling of northern Great Plains salt effleorescence
mineralogy; Soil Science Society of America Journal, v. 50,
p. 1363-1367.
1986b: Mineralogy and stability of soil evaporites in North Dakota; Soil
Science Society of America Journal, v. 50, p. 1069-1071.
Kelts, K. and Hsii, K.J.
1978: Freshwater carbonate sedimentation; in Lakes - Chemistry,
Geology, Physics, (ed.) A. Lerman; Springer Verlag, New York,
p. 295-324.
Kelts, K. and Talbot, M.R.
1989: Lacustrine carbonates as geochemical archives of environmental
change and biotic-abiotic interactions; in Ecological Structure and
Function of Large Lakes, (ed.) M.M. Tilzer and C. Seruya; Science
and Technology Publishers, Madison, Wisconsin, p. 290-317.
Kendall, A.C.
1984: Evaporites; in Facies Models, (ed.) R.G. Walker; Geoscience
Canada, Reprint Series 1, p. 259-296.
Kenny, B.C.
1985: Sediment resuspension and currents in Lake Manitoba; Journal of
Great Lakes Research, v. 11, p. 85-96.
Klassen, R.W.
1989: Quaternary geology of the southern Canadian Interior Plains; in
Chapter 2 of Quaternary Geology of Canada and Greenland, (ed.)
R.J. Fulton; Geological Survey of Canada, Geology of Canada,
no. 1, p. 138-173, (also Geological Society of America, The
Geology of North America, v. K-I).
1994: Late Wisconsinan and Holocene history of southwestern
Saskatchewan; Canadian Journal of Earth Sciences, v. 31,
p. 1822-1 937.
Kushnir, D.W.
1971: Sediments of the sonth basin of Lake Winnipeg; M.Sc. thesis,
University of Manitoba, Winnipeg, Manitoba, 92 p.
LaBaugh, J.W., Winter, T.C., Swanson, G.A., Rosenberry, D.O.,
Nelson, R.D., and Euliss, N.H.
1996: Changes in atmospheric circulation patterns affect midcontinent
wetlands sensitve to climate; Limnology and Oceanography,v. 41,
p. 864-870.
Lambert, S.
1989: Hydrogeochemistry of saline lakes in the northern Interior Plains of
western Canada and northern United States; B.Sc. thesis, University
of Manitoba, Winnipeg, Manitoba, 162 p.
Langham, E.J.
1970: Interim report and program review of the Big Quill Lake study;
Saskatchewan Research Council, Saskatoon, Saskatchewan,
Report E 70-12.39 p.
Last, W.M.
1982: Holocene carbonate sedimentation in Lake Manitoba, Canada;
Sedimentology, v. 29, p. 691-704.
1984a: Sedimentology of playa lakes of the northern Great Plains;
Canadian Journal of Earth Sciences, v. 21, p. 107-125.
1984b: Modern sedimentology and hydrology of Lake Manitoba, Canada;
Environmental Geology, v. 5, p. 177-190.
1987: Sedimentology,geochemistry, and evolution of a saline playa in the
northern Great Plains of Canada; in Geochemistry and Mineral
Formation in the Earth Surface, (ed.) R. Rodriguez-Clemente and
Y. Tardy; Consejo Superior de Investigaciones Cientificas, Centre
National dela RechercheScientifique,Madrid, Spain, p. 3 19-337.
1988: Salt lakes of western Canada: a spatial and temporal geochemical
perspective; in Proceedings of Symposium on Water Management
Affecting the Wet-to-Dry Transition, (ed.) W. Nicholaichuk and
H. Steppuhn; Water Studies Institute, Saskatoon, Saskatchwan,
p. 99-1 13.

Last, W.M. (cont.)


1989a: Continental brines and evaporites of the northern Great Plains of
Canada; Sedimentary Geology, v. 64, p. 207-221.
1989b: Sedimentology of a saline playa in the northern Great Plains,
Canada; Sedimentology, v. 36, p. 109-123.
1990: Lacustrine dolomite - an overview of modern, Holocene, and
Pleistoceneoccurrences;Earth-ScienceReviews, v. 27, p. 221-263.
1991a: Sedimentology, geochemistry, and evolution of saline lakes of the
Saskatoon area; Sedimentary and Paleolimnological Records of
Saline Lakes Conference, Saskatoon, August, 1991,
Within-ConferenceFieldtrip Guidebook, 98 p.
1991b: Sedimentology, geochemistry, and evolution of saline lakes of the
northern Great Plains; Sedimentary and PaleolimnologicalRecords
of Saline Lakes Conference, Saskatoon, August, 1991,
Post-ConferenceFieldtrip Guidebook, 175 p.
1992a: Chemical composition of saline and subsaline lakes of the northern
Great Plains, western Canada; International Journal of Salt Lake
Research, v. 1, p. 47-76.
1992b: Salt lake paleolimnology in the northern Great Plains: the facts, the
fears, the future; in Aquatic Ecosystems in Semi-Arid Regions,
(ed.) R.D. Robarts and M.L. Bothwell; Environment Canada,
Saskatoon, Saskatchewan, National Hydrology Research Institute,
Symposium Series 7, p. 51-62.
1993a: Geolimnologyof Freefight Lake: an unusual hypersaline lake in the
northern Great Plains of western Canada; Sedimentology, v. 40,
p. 431448.
1993b: Rates of sediment deposition in a hypersaline lake in the northern
Great Plains, western Canada; International Journal of Salt Lake
Research, v. 2, p. 47-58.
1994a: Paleohydrologyof playas in the northern Great Plains: perspectives
from Palliser's Triangle; in Paleoclilnate and Basin Evolution of
Playa Systems, (ed.) M.R. Rosen; Geological Society of America,
Special Paper 289, p. 69-80.
1994b: Deep-water evaporite mineral formation in lakes of western
Canada; in Sedirnentology and Geochemistry of Modern and
Ancient Saline Lakes, (ed.) R. Renaut and W.M. Last; Society of
Economic Paleontologists and Mineralogists, Special Publication
No. 50, p. 51-60.
1995: Evolution of saline lakes in western Canada; in Proceedings:
Climate, Landscape, and Vegetation Change in the Canadian
Prairie Provinces, (ed.) I.D. Campbell, C. Campbell, D.S. Lemmen,
and R.E. Vance; Canadian Forestry Service, Edmonton, Alberta,
p. 55-64.
1996a: Bulk composition, texture, and mineralogy of Lake Winnipeg core
and surface grab samples; in Lake Winnipeg Project: Cruise Report
and Scientific Results, (ed.) B.J. Todd, C.F.M. Lewis,
L.H. Thorleifson, and E. Nielsen; Geological Survey of Canada,
Open File 31 13, p. 209-220.
1996b: Geochemistry, texture, and mineralogy of Lake Winnipeg sediments; in Lake Winnipeg Project: Cruise Report and Scientific
Results, (ed.) B.J. Todd, C.F.M. Lewis, L.H. Thorleifson, and
E. Nielsen; Geological Survey of Canada, Open File 3113,
p. 485-584.
1996c: Lithostratigraphy and mineralogy of Oro Lake, Saskatchewan;
Canadian Association of Geographers Annual Meeting, University
of Saskatoon, Saskatoon, Saskatchewan, Program and Abstracts,
p. 130-131.
Last, W.M. and Gosselin, D.C. (ed.)
1997: Paleolimnology in the Great Plains of North America - special
issue; Journal of Paleolimnology, v. 17, p. 1-154.
Last, W.M. and Sauchyn, D.J.
1993: Mineralogy and lithostratigraphy of Harris Lake, sor~thwestern
Saskatchewan;Canadian Journalof Paleolimnology,v. 9, p. 23-39.
Last, W.M. and Schweyen, T.H.
1983: Sedimentology and geochemistry of saline lakes of the northern
Great Plains; Hydrobiologia, v. 105, p. 245-263.
1985: Late Holocene history of Waldsea Lake, Saskatchewan, Canada;
Quaternary Research, v. 24, p. 219-234.
Last, W.M. and Slezak, L.A.
1986: Paleohydrology, sedimentology, and geochemistry of two deep
saline lakes from southern Saskatchewan; GBographie physique et
Quaternaire, v. 1 1, p. 5-15.
1987a: Sodium sulfate deposits of western Canada; in Economic Minerals
of Saskatchewan, (ed.) C.F. Gilboy and L.W. Vigrass;
Saskatchewan Geological Society, Regina, Saskatchewan, Special
Publication No. 8, p. 197-205.

GSC Bulletin 534


Last, W.M. and Slezak, L.A. (cont.)
1987b: Geolimnology of Freefight Lake, Saskatchewan; irz SLEADS
Workshop 87, (ed.) A.R. Chivas and P. De Deckke~;The Australian
National University, Canberra, Australia, p. 9-1 1.
1988: The salt lakes of western Canada: a paleolimnological overview;
Hydrobiologia, v. 158, p. 301-316.
Last, W.M. and Teller, J.T.
1997: Paleolimnology of Lake Manitoba: solutions, complicatio~is,and
contradictions in the record of a large lake in the Prairies of western
Canada; Wiirzburger Geographische Manuskripte, Heft 41,
p. 117-118.
Last, W.M. and Vance, R.E.
1996: Stop 12: Antelope Lake; in Landscapes of the Palliser Triangle,
Guidebook for the Canadian Geomorphology Research Group
Field Trip, (ed.) D.S. Lemmen; Canadian Association of
Geographers, 1996 Annual Meeting, Saskatoon, Saskatchewan,
p. 42-43.
1997: Bedding characteristics of Holocene sediments from salt lakes of
the northern Great Plains, western Canada; Journal of
Paleolimnology, v. 17, p. 297-318.
Last, W.M., Teller, J.T., and Forester, R.M.
1994: Paleohydrology and paleochernistry of Lake Manitoba, Canada: the
isotope and ostracode records; Journal of Paleolimnology, v. 12,
p. 269-282.
Leckie, D.A. and Smith, D.G.
1993: Regional setting, evolution and depositional cycles of the western
Canadian foreland basin; in Foreland Basins, (ed.) R. Macqueen
and D.A. Leckie; American Association of Petroleum Geologists,
Memoir 55, p. 9 4 6 .
Lemmen, D.S., Dyke, L.D., and Edlund, S.A.
1993: The Geological Survey of Canada's Integrated Research and
Monitoring Area (IRMA) projects: a contribution to Canadian
global change research; Journal of Paleolimnology,v. 9, p. 77-83.
Lemmen, D.S., Vance, R.E., Campbell, I.A., David, P.P.,
Pennock, D.J., Sauchyn, D.J., and Wolfe, S.A.
1998: Geomorphic systems of the Palliser Triangle, southern Canadian
Prairies: description and response to changing climate; Geological
Survey of Canada, Bulletin 521,72 p.
Lemmen, D.S., Vance, R.E., Wolfe, S.A., and Last, W.M.
1997: Impacts of future climate change on the southern Canadian Prairies:
a paleoenvironrnental perspective; Geoscience Canada. v. 24,
p. 121-133.
Lennox, D.H., Maathuis, H., and Pederson, D.
1988: Region 13, western glaciated plains; irz Hydrogeology, (ed.)
W. Back, J.S. Rosenshein, and P.R. Seaber; Geological Society of
America, The Geology of North America, v. 0-2, p. 115-128.
Lieffers, V.J.
1981: Environment and ecology of Sciipus rnariti7nus L. var. apludosus
(nels.) Kuk. in saline wetlands of the Canadian Prairies; Ph.D.
thesis, University of Manitoba, Winnipeg, Manitoba, 190 p.
Lieffers, V.J. and Shay, J.M.
1983: Ephemeral saline lakes in the Canadian Prairies - their classification and management for emergent macrophyte growth;
Hydrobiologia, v. 105, p. 85-94.
Lockhart, E.B.
1984: Origin of gypsum in Lydden Lake, Saskatchewan; B.Sc. thesis,
University of Manitoba, Winnipeg, Manitoba, 57 p.
Lockhart, E.B. and Last, W.M.
1984: Sedimentology and geochemistry of a small salt playa in southern
Saskatchewan; Geological Association of Canada, Program with
Abstracts, v. 9, p. 84.
Lowenstein, T.K. and Hardie, L.A.
1985: Criteria for the recognition of salt-pan evaporites; Sedi~nentology,
v. 32, p. 627-644.
MacWilliams, R.W. and Reynolds, R.G.
1973: Solution mining of sodium sulfate; Canadian Institute of Mining
and Metallurgy Transactions, v. LXXVI, p. 127-131.
Meneley, W.A., Christiansen, E.A., and Kupsch, W.O.
1957: Preglacial Missouri River in Saskatchewan; Journal of Geology,
v. 65, p. 441-447.
Mills, G.F. and Smith, R.E.
1971: Soils of the Grahamdale area; Manitoba Soil Survey, Soil Report
No. 16, 132 p.
Mitchell, P. and Prepas, E. (ed.)
1990: Atlas of Alberta Lakes; University of Alberta Press, Edmonton,
Alberta, 675 p.

Mollard, J.D. and Janes, J.R.


1984: Airphoto interpretation and the Canadian landscape; Energy,
Mines, and Resources Canada, Ottawa, 415 p.
Mollard, J.D. and Mollard, D.G.
1987: Location, exploration and production of gravel resources in
Saskatchewan; in Economic Minerals of Saskatchewan, (ed.)
C.F. Gilboy and L.W. Vigrass; Saskatchewan Geological Society,
Regina, Saskatchewan, Special Publication No. 8, p. 21 1-216.
Moller, P. and Kubanek, F.
1976: Role of magnesium in nucleation processes of calcite, aragonite and
dolomite; Neues Jahrbuch fiir Mineralogie Abhandlungen, v. 126,
p. 199-220.
Moore, G.E.
1939: A limnological study of certain lakes of southern Saskatchewan;
M.A. thesis, University of Saskatchewan, Saskatoon,
Saskatchewan,40 p.
Moorhouse, W.W.
1959: The Study of Rocks in Thin Section; Harper & Brothers, New York,
514 p.
Moran, S.R., Groenewold, G.H., and Cherry, J.A.
1978: Geologic, hydrologic, and geochemical concepts and techniques in
characterization for mined land reclamation; North Dakota
Geological Survey, Report of Investigations No. 63, 152 p.
Mossop, G. and Shetsen, I. (coinp.)
1994: Geological Atlas of Western Canada Sedimentary Basin; Canadian
Society of Petroleum Geologists and Alberta Research Council,
Calgary, Alberta, 510 p.
Miiller, G., Irion, G., and Forstner U.
1972: Formation and diagenesis of inorganic Ca-Mg-carbonates in the
lacustrine environment; Natunvissenschaften, v. 59, p. 158-164.
Murphy, E.C.
1996: The sodium sulfate deposits of northwestern North Dakota; North
Dakota Geological Survey, Report of Investigations No. 99.73 p.
Neal, J.T.
1975: Introduction;irzPlayasandDried Lakes, (ed.) J.T. Neal; Benchmark
Papers in Geology, Dowden, Hutchinson, and Ross Inc.,
Stroudsburg, Pennsylvania, p. 1-5.
Northcote, T.G. and Larkin, P.A.
1963: Chapter 16: western Canada; irz Limnology in North America, (ed.)
D.G. Frey; University of Wisconsin Press, Madison, Wisconsin,
p. 451485.
Oldfield, F.
1996: PALICLAS: the partnership, its perspectives and goals; in
Palaeoenvironmental Analysis of Italian Crater Lake and Adriatic
sediments, (ed.) P. Guilizzoni and F. Oldfield; Memorie
dell'lstituto Italiano di Idrobiologia, v. 55, p. 7-16.
Osterkamp, W.R. and Wood, W.W.
1987: Playa-lake basins on the southern High Plains of Texas and New
Mexico: part I, hydrologic, geomorphic, and geologic evidence for
their development; Geological Society of America Bulletin, v. 99,
p. 215-223.
Owen, J.B., Tubb, R., Anderson, D.W., and Callender, E.
1973: The biogeochemistry of Devils Lake, North Dakota; United States
Office of Water Resources and Research (OWRR), Project Report
A-006-NDAK, 83 p.
Padden, M.C.
1996: Holocene paleohydrology of the Palliser Triangle from isotope
studies of lake sediments; M.Sc. thesis, University of Waterloo,
Waterloo, Ontario, 108 p.
Palliser, J.
1862: Journals, detailed reports and observations relative to the exploration by Captain Palliser of a portion of British North America;
Exploration of British North America: Report Presented to Both
Houses of Parliament, London, United Kingdom, 325 p.
Pearson, F.J. and Coplen, T.B.
1978: Stable isotope studies of lakes; in Lakes - Chemistry, Geology,
Physics, (ed.) A. Lerman; Springer Verlag, New York, p. 325-339.
Perkins, R.D., Dwyer, G.S., Rosoff, D.B., Fuller, J., Baker, P.A.,
and Lloyd, R.M.
1994: Salina sedimentation and diagenesis: West Caicos Island, British
West Indies; in Dolomites, (ed.) B. Pruser, M. Tucker, and
D. Zenger; International Association of Sedimentologists, Special
Publication No. 21, p. 55-74.
Peryt, T.M. (ed.)
1987: Evaporite basins; Lecture Notes in Earth Sciences 13,
Springer-Verlag,New York, 188 p.

W.M. Last
Pupp, C., Maathuis, H., and Grove, G.
1991: Groundwater quality in Saskatchewan: hydrogeology, quality concerns, management; National Hydrology Research Institute
Contribution No. CS-91028.61 p.
Purser, B., Tucker, M., and Zenger, D.
1994: Dolomites; International Association of Sedimentologists,Special
Publication No. 21,450 p.
Rawson, D.S. and Moore, G.E.
1944: The saline Lakes of Saskatchewan; Canadian Journal of Research
(Series D), v. 22, p. 141-201.
Reeves, C.C.
1968: Introduction to Paleolimnology; Elsevier, Amsterdam,
Netherlands, 228 p.
Reeves, C.C. (ed.)
1972: Playa lake symposium;International Center for Arid and Semi-Arid
Land Studies, Lubbock, Texas, 334 p.
Remenda, V.H. and Birlts, S.J.
1999: Groundwater in the Palliser Triangle: an overview of its vulnerability and potential to archive climate information; in Holocene
Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Remenda, V.H., Cherry, J.A., and Edwards, T.W.D.
1994: Isotopic compositionof old groundwater fromLake Agassiz: implications forlatePleistoceneclimate; Science, v. 266, p. 1975-1978.
Renaut, R.W. and Last, W.M.
1994: Salt lake sedimentology and geochemistry:introduction and historical perspective; in Sedimentology and Geochemistry of Modern
and Ancient Saline Lakes, (ed.) R.W. Renaut and W.M. Last;
Society of Econo~nicPaleontologists and Mineralogists, Special
Publication No. 50, p. v-x.
Renaut, R.W. and Long, P.R.
1989: Sedimentology of saline lakes of the Cariboo Plateau, interior
British Columbia; Sedimentary Geology, v. 64, p. 239-264.
Richardson, J.L.
1969: Former lake-level fluctuations their recognition and interpretation. Mitteilungen des InternationalenVereinigung fiir Limnologie,
v. 17, p. 78-93.
Roberts, S.M. and Spencer, R.J.
1995: Paleotemperatures preserved in fluid inclusions in halite;
Geochimica et Cosmochimica Acta, v. 59, p. 3929-3942.
Roberts, S.M., Spencer, R.J., Yang, W., and Krouse, H.R.
1997: Deciphering some unique paleotemperature indicators in
halite-bearing saline lake deposits from Death Valley, California,
U.S.A.; Journal of Paleolimnology, v. 17, p. 101-130.
Rosen, M.R.
1994: The importanceof groundwaterin playas: a review of playa classifications and the sedimentology and hydrology of playas; in.
Paleoclimate and Basin Evolution of Playa Systems, (ed.)
M.R. Rosen; Geological Society of America, Special Paper 289,
p. 1-18.
Rosen, M.R., Miser, D.E., Starcher, M.A., and Warren, J.K.
1989: Formation of dolomite in the Coorong region, South Australia;
Geochemica et Cosmochimica Acta, v. 53, p. 661469.
Rbikowska, A.D. and Rbikowski, A.
1969: Seasonal changes of slough and lake water chemistry in southern
Saskatchewan,Canada; Journal of Hydrology, v. 7, p. 1-13.
R6iltowski, A.
1967: The origin of hydrochemical patterns in hummocky moraine;
Canadian Journal of Earth Sciences, v. 4, p. 1065-1092.
Rueffel, P.G.
1968a: Natural sodium sulfate in North America; in Third Symposium on
Salt, (ed.) J.L. Rau and L.F. Dellwig; Northern Ohio Geological
Society, Cleveland, Ohio, p. 429-451.
1968b: Development of the largest sodium sulfate deposit in Canada;
Canadian Institute of Mining amid Metallurgy Bulletin, v. 61,
p. 1217-1228.
Rutherford, A.A.
1967: Water quality survey of Saskatchewan groundwaters;
Saskatchewan Research Council, Saskatoon, Saskatchewan,
Report C-66-1,267 p.
1970: Water quality survey of Saskatchewan surface waters;
Saskatchewan Research Council, Saskatoon, Saskatchewan,
Report C-70-1, 133 p.

Sack, L.A.
1993: Geolimnology and paleolimnology of Little Manitou Lake,
Saskatchewan; B.Sc. thesis, University of Manitoba, Winnipeg,
Manitoba, 52 p.
Sack, L.A. and Last, W.M.
1994: Lithostratigraphy and recent sedimentation history of Little
Manitou Lake, Saskatchewan,Canada; Journal of Paleolimnology,
V. 10, p. 199-212.
Sahinen, U.M.
1948: Preliminary report on sodium sulfate in Montana; Montana Bureau
of Mines and Geology, Information Circular No. 11.9 p.
Sauchyn, D.J.
1997: Proxy records of postglacial climate in the Canadian Prairie Provinces: a guide to literature and current research; in Responding to
Global Climate Change in the Prairies, Volume 111 of the Canada
Country Study: Clirnate Impacts and Adaptation, (ed.)
R. Herrington, B. Johnson, and F. Hunter; Environment Canada,
Prairie and Northern Region, Appendix I, p. 1-44.
Schwartz, F.W. and Gallup, D.N.
1978: Some factors controlling the major-ion chemistry of small lakes;
examples from the prairie parkland of Canada; Hydrobiologia,
v. 58, p. 65-81.
Schweyen, T.H.
1984: Sedimentology and paleohydrology of Waldsea lake,
Saskatchewan, an ectogenic mecomictic saline lake; M.Sc. thesis,
University of Manitoba, Winnipeg, Manitoba, 111 p.
Schweyen, T.H. and Last, W.M.
1983: Sedimentology and paleohydrology of Waldsea Lake,
Saskatchewan; in Third Biennial Plains Aquatic Research Conference: Proceedings, (ed.) M.D. Scott; Canadian Plains Proceedings,
v. l I, p. 45-59.
Scott, M.D. and Scott, S.A.
1986: Saline lake management for waterfowl resources in western North
America: a review of concepts; in Evaluating Saline Waters in a
Plains Environment, (ed.) D.T. Waite; Canadian Plains Research
Centre, University of Regina, Regina, Saskatchewan, Canadian
Plains Proceedings, v. 17, p. 23-38.
Shang, Y. and Last, W.M.
1999: Mineralogy, lithostratigraphy, and inferred geochemical history of
North Ingebrigt lake, Saskatchewan; in Holocene Climate and
Environmental Change in the Palliser Triangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.
Shukla, V. and Baker, P.A. (ed.)
1988: Sedilnentology and geochemistry of dolostones; Society of
Economic Paleontologists and Mineralogists, Special Publication
No. 43,266 p.
Sltarie, R.L., Richardson, J.L., McCarthy, G.J., and Maianu, A.
1987: Evaporite mineralogy and groundwater chemistry associated with
saline soils in eastern North Dakota; Soil Science Society of
America Journal, v. 51, p. 1272-1377.
Slezak, L.A.
1989: Sedimentology of Freefight Lake, Saskatchewan; M.Sc. thesis,
University of Manitoba, Winnipeg, Manitoba, 120 p.
Slezak, L.A. and Last, W.M.
1985: Geology of sodium sulfate deposits of the northern Great Plains; in
Twentieth Forum on the Geology of Industrial Minerals, (ed.)
J.D. Glaser and J. Edwards; Maryland Geological Survey, Special
Publication No. 2, p. 105-1 15.
Sloan, C.E.
1972: Groundwaterhydrology of prairie potholes in North Dakota; United
States Geological Survey, Professional Paper 585C, 28 p.
Sly, P.G.
1978: Sedimentaly processes in lakes; i ~ zLakes - Chemistry, Geology,
Physics, (ed.) A. Lerman; Springer Verlag, New York, p. 65-90.
Smoot, J.P. and Lowenstein, T.K.
1991: Depositional environments of non-marine evaporites; irz
Evaporites, Petroleum and Mineral Resources, (ed.) J.L. Melvin;
Elsevier, New York, p. 189-348.
Sonnenfeld, P.
1984: Brines and Evaporites; Academic Press, New York, 613 p.
Sonnenfeld, P. and Perthuisot, J.P.
1989: Brines and evaporites; American Geophysical Union, Washington,
D.C., Short Course in Geology, v. 3, 126 p.

GSC Bulletin 534


Stalker, A.
1961: Buried valleys in central and southern Alberta; Geological Survey
of Canada, Paper 60-32, 13 p.
Swart, P.K., Lohmann, K.C., McKenzie, J., and Savin, S.
1993: Climatic change in continental isotopic records; American
Geophysical Union, Geophysical Monograph 78, Washington,
D.C., 374 p.
Talbot, M.R. and Allen, P.A.
1996: Lakes; in Sedimentary Environments: Processes, Facies and
Stratigraphy, (ed.) H.G. Reading; Blackwell Science Ltd, Oxford,
United Kingdom, p. 83-124.
Talbot, M.R. and Kelts, K.
1986: Primary and diagenetic carbonates in the anoxic sediments of Lake
Bosumtwi, Ghana; Geology, v. 14, p. 912-916.
Teller, J.T. and Kehew, A.E. (ed.)
1994: Late glacial history of large proglacial lakes and meltwater lunoff
along the Laurentide Ice Sheet; Quaternary Science Reviews, v. 13,
p. 795-981.
Teller, J.T. and Last, W.M.
1979: Post-glacialsedimentation and history of Lake Manitoba; Manitoba
Mines, Natural Resources, and Environmental Management,
Report No. 79-41, 185 p.
1981: Late Quaterna~yhistoly of Lake Manitoba, Canada; Quaternary
Research, v. 16, p. 97-1 16.
1990: Paleohydrological indicators in playasand saltlakes, with examples
from Canada, Australia, and Africa; Palaeogeography,
Palaeoclimatology,Palaeoecology,v. 76, p. 215-240.
Teller, J.T., Thorleifson, L.H., Matile, G., Brisbin, W.C.,
and Nielsen, E.
1996: Sedimentology, geomorphology, and history of the central Lake
Agassiz basin; Geological Association of Canada, AnnualMeeting,
Winnipeg, Manitoba, Fieldtrip B2 Guidebook, 76 p.
Thomas, J.F.J.
1959: Industrial water resources of Canada. Nelson River Drainage Basin
in Canada, 1953-1956. Canadian Water Survey, Report No. 10,
147 p.
Thompson, R. and Oldfield, F.
1986: Environmental Magnetism; Allen & Unwin, London, United
Kingdom, 227 p.
Timms, B.V.
1992: Lake Geomorphlogy; Gleneagles Publishing, Adelaide, Australia,
180 p.
Timpson, M.E. and Richardson, J.L.
1986: Ionic composition and distribution ion saline seeps of southwestern
North Dakota; Geoderma, v. 37, p. 295-305.
Timpson, M.E., Richardson, J.L., Keller, L.P., McCarthy, G.J.
1986: Evaporite mineralogy associated with saline seeps in southwestern
North Dakota; Soil Science of America Journal, v. 50, p. 490-493.
Todd, B.J., Lewis, C.F.M., Thorleifson, L.H., and Nielsen, E.
1996: Lake Winnipeg Project: cruise report and scientific results;
Geological Survey of Canada, Open File 31 13,656 p.
Todd, B.J., Lewis, C.F.M., Thorleifson, L.H., Nielsen, E.,
and Last, W.M.
1998: Paleolimnology of Lake Winnipeg: introduction and overview;
Journal of Paleolimnology, v. 19, p. 21 1-213.
Tokarsky, 0.
1985: Hydrogeologic profile, Alberta-Saskatchewan boundary; Prairie
Province Water Board, 63 p.
1986: Hydrogeologicprofile, Saskatchewan-Manitoba boundary; Prairie
Province Water Board, Report No. 79,45 p.
Tomkins, R.V.
1954a: Natural sodium sulfate in Saskatchewan; Saskatchewan Department of Mineral Resources, Report 6,71 p. (second edition).
1954b: Magnesium in Saskatchewan; Saskatchewan Department of Mineral Resources, Report l l , 2 3 p.
Tremblay, J.Y.
1984: Industrial minerals; Canadian Mining Journal, v. 105, p. 85-87.
Vance, R.E.
1991: A paleobotanical study of Holocene drought frequency in southern
Alberta; Ph.D. thesis, Simon Fraser University, Burnaby, British
Columbia, 180 p.
1997: The Geological Survey of Canada's Palliser Triangle Global
Change Project: a multidisciplinary geolimnological approach to
predicting potential global change impacts on the northern Great
Plains; Journal of Paleolimnology, v. 17, p. 3-8.

Vance, R.E. and Last, W.M.


1994: Paleolimnology and global change on the southern Canadian
Prairies; in Current Research 1994-B; Geological Survey of
Canada, p. 49-58.
1996: Stop 4: Oro Lake; in Landscapes of the Palliser Triangle, Guidebook for the Canadian Geomorphology Research Group Field Trip,
(ed.) D.S. Lemmen; Canadian Association of Geographers, 1996
Annual Meeting, Saskatoon, Saskatchewan, p. 26-27.
Vance, R.E., Beaudoin, A.B., and Luckman, B.H.
1995: The paleoecological record of 6 ka BP climate in the Canadian
Prairie Provinces; Geographie physique et Quaternaire, v. 49,
p. 81-98.
Vance, R.E., Clague, J.J., and Mathewes, R.W.
1993: Holocene paleohydrology of a hypersaline lake in southeastern
Alberta; Journal of Paleolimnology, v. 8, p. 103-120.
Vance, R.E., Last, W.M., and Lemmen, D.S.
1998: The rain (or the lack therof) on the plain: putting the Canadian Great
Plains on the paleoclimate map; in Past Global Changes and the
Significance for the Future; PAGES Open Science Meeting,
University of London, London, United Kingdom, p. 128.
Vance, R.E., Last, W.M., and Smith, A.J.
1997: Hydrologic and climatic implications of a multidisciplinary study
of late Holocene sediment from Kenosee Lake, southeastern
Saskatchewan, Canada; Journal of Paleolimnology, v. 18,
p. 365-393.
Vance, R.E., Mathewes, R.W., and Clague, J.J.
1992: 7000 year record of lake-level change on the northern Great Plains:
a high-resolution record proxy of past climate; Geology, v. 20,
p. 879-882.
van Everdingen, R.O.
1971: Surface-water composition in southern Manitoba reflecting discharge of saline subsurface waters and subsurface solution of
evaporites; irt Geoscience Studies in Manitoba, (ed.) A.C. Turnock;
Geological Association of Canada, Special Paper 9, p. 343-352.
Van Stempvoort, D.R., Edwards, T.W.D., Evans, M.S.,
and Last, W.M.
1993: Paleohydrology and paleoclimate records in a saline prairie lake
core: mineral, isotope and organic indicators; Journal of
Paleolimnology, v. 8, p. 135-147.
Veldman, W.M.
1969: Shoreline processes on Lake Winnipeg; M.Sc. thesis, University of
Manitoba, Winnipeg, Manitoba, 82 p.
Waite, D.T. (ed.)
1986: Evaluating saline waters in a plains environment; Canadian Plains
Research Centre, University of Regina, Regina, Saskatchewan,
Canadian Plains Proceedings, v. 17, 107 p.
Wallick, E.I.
1981: Chemical evolution of groundwaterin a drainage basin of Holocene
age, east-central Alberta, Canada; Journal of Hydrology, v. 54,
p. 245-283.
Wallick, E.I. and grouse, H.R.
1977: Sulfur isotope geochemistry of a groundwater-generated
Na2S04NazC03 deposit and the associated drainage basin of
Horseshoe Lake, Metiskow, east central Alberta, Canada; Second
International Symposium on Water-rock Interaction, Strasbourg,
France, p. 1156-1 164.
Warren, J.K.
1989: Evaporite Sedimentology: Importance in Hydrocarbon Accumulation; Prentice-Hall, Inc., Englewood Cliffs, New Jersey, 285 p.
Weisman, W.I. and Tandy, C.D.
1975: Sodium sulphate deposits; in Industrial Minerals and Rocks, (ed.)
S.J. Lefond; American Institute of Mining, Metallurgical and
Petroleum Engineers, New York, p. 1081-1093 (fourth edition).
Whiting, J.M.
1974: Closed drainage lakes are a climatic change indicator;
Saskatchewan Research Council, Saskatoon, Saskatchewan,
Report E74- 10.3 1 p.
1977: The hydrologic and chemical balance of the Big Quill Lake basin;
Saskatchewan Research Council, Saskatoon, Saskatchewan,Publication E77-12, 98 p.
1986: The effects of climate on prairie lakes; in Proceedings of the
Canadian Hydrology Symposium No. 16, Drought: The Impending
Crisis? Associate Committee on Hydrology, Regina,
Saskatchewan, p. 639456.
Wilby, T.
1914: A Motor Tour Through Canada; Bell & Cockburn,Toronto, 290p.

W.M. Last

Williams, W.D.
1967: The chemical characteristicsof lentic surface waters in Australia; in
Australian lnland Waters alld Their Fauna, (ed.) A.H. Weatherley;
Australian National University Press, Canberra, Australia,
p. 18-77.
Williams, W.D. and Sherwood, J.E.
1994: Definition and measurement of salinity in salt lakes; International
Journal of Salt Lake Research, v. 3, p. 53-63.
Wilson, M.C. and Dijks, I.J.
1993: Introductory essay -land of no quarter, the Palliser Triangle as an
environmental-culturalpump; inThe Palliser Triangle, A Region in
Space and Time, (ed.) R.W. Barendrgt, M.C. Wilson, and
F.J. Jankunis; University of Lethbridge, Lethbridge, Alberta,
p. 37-61.
Winter. T.C.
1977: ' Classificationof the hydrologic settings of lakes in the north central
United States: Water Resources Research. v. 13. o. 753-767.
1978: Numerical simulation of steady-state &ee diGnsional groundwater flow near lakes; Water Resources Research, v. 14,
p. 245-254.
1995: Hydrological processes and the water budget of lakes; in Lakes Chemistry, Geology, Physics, (ed.) A. Lerman; Springer Verlag,
New York, p. 37-62.

Witkind, I.J.
1952: The localization of sodium sulfate deposits in northeasternMontana
and northwestern North Dakota: American Journal of Science.
v. 250, p. 667-676.
Wolfe, B.B. and Edwards, T.W.D.
1997: Hydrologic control on the oxygen-isotope relation between sediment cellulose and lake water, western Taimyr Peninsula, Russia:
implications for the use of surface-sediment calibrations in
paleolimnology; Journal of Paleolimnology, v. 18, p. 283-291.
Wolfe, B.B., Edwards, T.W.D., Aravena, R., and MacDonald, G.M.
1996: Rapid Holocene hydrologic change along boreal treeline revealed
by 613C and 6180 in organic lake sediments, Northwest Territories,
Canada; Journal of Paleolimnology, v. 15, p. 171-181.
Wolfe, S.A. and Lemmen, D.S.
1999: Monitoring of sand dune activity in the Great Sand Hills region,
southwestern Saskatchewan; i n Holocene Climate and
Environmental Change in the Palliser Triangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.
Wood, W.W. and Sanford, W.E.
1990: Groundwater control of evaporite deposition; Economic Geology,
v. 85, p. 1226-1235.
1995: Eolian transport, saline lake basins, andgroundwater solutes; Water
Resources Research, v. 31, p. 3121-3129.
Zenger, D.H. and Mazzullo, S.J. (ed.)
1982: Dolornitization; Hutchinson Ross Publishing Company,
Stroudsburg, Pennsylvania, 426 p.

GSC Bulletin 534

APPENDIX A
Economic geology
Lakes of the northern Great Plains are a source of valuable industrial materials, minerals, and compounds.
Exploitation of these resources in western Canada started well before the arrival of Europeans, evidenced by
journals and diaries of nineteenth century European settlers that commonly refer to Aboriginal use of the
lacustrine salts and brines for medicinal purposes, tanning, and food preservation. These salts also provided
the basis for several of the earliest commercial industrial efforts on the northern Great Plains (Hind, 1861;
Cole, 1930; Binyon, 1952). Large-scale mineral production from the lakes began in 1918 with the extraction
of magnesium and sodium sulphates and carbonates from Muskiki Lake near Saskatoon (Cole, 1926). Production of anhydrous sodium sulphate (salt cake) from some 20 different lakes gradually increased over the
next five decades to a high of approximately 700 000 tin 1973. Today, the region supplies nearly 50% of the
total North American demand for sodium sulphate. A large increase in the price of salt cake during
1973-1975 (from $15 to $48 per t) and again during 1980-1983 (from $62 to $108 per t) saw renewed
interest in leasing and mining activities in the region during these periods. Despite softening markets and
production declines during the last several years, price stabilization at about $90 pert has led to an average of
more than $40 000 000 of sodium sulphate produced annually from the lakes.
Historically, the two largest uses of sodium sulphate have been in the production of h a f t paper and allied
products, and in the manufacture of detergents (Broughton, 1984; Murphy, 1996). More recently, however,
the energy industry has been consuming larger amounts of the salt as a conditioner to facilitate fly-ash suppression in coal-burning power plants (Tremblay, 1984). Another new use of salt cake is in the manufacture
of potassium sulphate by the reaction of Na2S04 with KC1 (Barry et al., 1985; Barry, 1986; Eatock, 1987).
Other potentially significant applications include use in glass, ceramic, and paint manufacture, and in solar
energy collectors.
The sodium sulphate industry is based on reserves of three basic types: 1)the sodium sulphate that is dissolved in the lake water; 2) the hydrated sodium sulphate mineral mirabilite (Na2S04.10H20), which occurs
seasonally on the floor of the salt lakes due to precipitation within the overlying water column; and 3) the
bedded salts composed mainly of mirabilite and thenardite (Na2S04) that make up the Holocene sedimentary fill in some basins. Although each of these sources has been exploited during the past eighty years, most
production today is from the hypersaline lake waters which are pumped from the basin into holding reservoirs. Upon further concentration by evaporation during summer and then cooling during the fall season,
mirabilite (known to the miners as Glauber's salt) is precipitated from the solution. The overlying brine is
then drained back into the lake basin, and the salt is removed to stockpiles. Solution mining using hot water
(MacWilliams and Reynolds, 1973) and dredge mining of the permanent salt beds have also been used.
This harvested Glauber's salt must be dehydrated prior to marketing. The methods for this processing
vary considerably and are reviewed in detail by Weisman and Tandy (1975), Broughton (1984), and Rueffel
(1968a). Most producers simply raise the temperature of the salt above its fusion point (about 32OC) and then
either continue heating to evaporate the water of crystallization or remove the solid anhydrous precipitate
from the slurry.
In addition to sodium sulphate, some of the lakes contain marketable amounts of magnesium (in the form
of both magnesium sulphates and carbonates; Tomkins (1954b)), sodium bicarbonate (baking soda; Cole
(1926), C.W. Templeton (Geological Study and Reserves Report, Metiskow Lakes Salt Deposit,
Township 39 Range 6W4, Alberta, unpub. report prepared for Canadian Chemical Company, 1968),
Alberta Sulphate Limited (Evaluation of remaining salt deposits, unpub. report, 1981)), and sodium chloride (Cole, 1930). Finally, coarse clastic sediments (sand, gravel) deposited on beaches, along shorelines,
and in deltas of both proglacial and modern lakes are utilized by many of the urban communities in the
region (James F. MacLaren Limited, 1980; Groom, 1985; Mollard and Mollard, 1987).

W.M. Last

APPENDIX B
Measurement of water salinity
Many methods are used to measure water salinity (Williams and Sherwood, 1994), so that comparison of
data sets can often be difficult and potentially misleading. Biological limnologists and ecologists often use
conductivity as a measure of salinity. Conductivity, or specific conductance, is a measure of the ease with
which electrical current will pass through the water. In general, the greater the salinity, the greater the conductivity. However, this relationship, which is controlled by the specific ions present in the solution as well
as the level of concentration of the ions, is not straightforward. For example, a conductivity of 126mS2,, in a
lake such as Bitter Lake in southwestern Saskatchewan, dominated by sodium and chloride ions, would be
equivalent to about 100 parts per thousand (ppt or %o) total dissolved solids (TDS), but this same conductivity would also be recorded in a brine having only about 75 ppt TDS that was dominated by magnesium and
sulphate ions (Desai and Moore, 1969). Clearly, the use of conductivity to estimate salinity in lakes having
diverse chemical compositions, such as these basins in western Canada, should be avoided.
The quantity or abundance of individual ionic components in the water is usually reported as weight concentration of the ion in solution (i.e. the weight of the dissolved ion in grams or milligrams per kilogram of
solution). This is preferred to the use of weight per litre because of the large increase in density of the solution at high salinities. Care must be exercised in evaluating and comparing published reports using g . ~ - l
units; unless density is taken into account by the analyst, the difference can be substantial at elevated salinities (>7 ppt TDS; Hem (1985)). For example, the bottom water of Freefight Lake, a deep meromictic lake in
Saskatchewan, is 275 ppt (weightlweight) or 340 g.L-I. It follows then that the total salinity (total dissolved
solids) should be the sum of the measured ionic components.
Those involved with appraising lake waters from a physical chemistry perspective, and particularly with
examining the thermodynamics of the aqueous/mineral reactions occurring in lakes, usually report water
composition using traditional chemical concentration nomenclature: molal units (m; moles per kilogram of
solvent) or molar units (M; moles per litre of solution) and in equivalents. Equivalent, equivalent weight, or
combining weight is the formula weight of the dissolved component divided by its valence. Equivalent
weight units are particularly useful when comparing ionic ratios of one water with another.
In fresh-water systems, molal and molar units are essentially the same. However, in concentrated solutions, such as those in the lakes of western Canada, differences become significant. Similarly, discussions
involving mineral precipitation/dissolution in the lakes usually evaluate thermodynamic activity of the ions
( a ) and ionic strength (I;one half the sum of the product of molality times the square of the valence of each
ion). Again, in waters having less than about 30 ppt TDS, there is relatively little difference between thermodynamic activity and laboratory-derived molality. At higher concentrations, however, the electrostatic
interactions between the ions greatly reduce their thermodynamic 'concentration' and a is considerably
smaller than m. Like conductivity, ionic strength is also highly dependent on the specific ions in solution: a
1 m NaCl solution has an ionic strength of 1, whereas the I of a solution having the same analytical concentration but dominated by sodium and sulphate is 3 (Drever, 1988).

Groundwater in the Palliser Triangle:


an overview of its vulnerability and
potential to archive climate information
V.H. ~ernenda'and S.J.

irks^

Remenda, V.H. and Birks, S.J., 1999: Groundwater in the Palliser Triangle: an overview of its vulnerability and potential to archive climate information; in Holocene Climate and Environmental
Change in the Palliser Triangle: A Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemrnen and R.E. Vance; Geological
Survey of Canada, Bulletin 534, p. 57-66.

Abstract: Groundwater is an important but poorly delineated resource in the drought-sensitive Palliser
Triangle. Much of the rural population of the region relies on groundwater as a water supply. Groundwater
discharge provides base flow to many surface-water bodies, including shallow lakes. Both bedrock and glacial deposits form important aquifers, although their water quality and sensitivity to climate fluctuations,
over exploitation, and contamination v a y . Ironically, it is the most vulnerable aquifers that contain the highest quality water. Groundwater is a potential archive of past climate fluctuations, both directly in the form of
stable isotope concentrations, and indirectly in the depths to which water tables declined in response to
Holocene climate change. To date, no observations of such changes have been made at temporal scales useful for drought management. The spatial and temporal scales of groundwater systems discharging to lakes,
the hydrochemistry, and the volume and source of water all have the potential to damp the connection
between proxy indicators of lake level and precipitation.
Rbsumb : Dans le Triangle de Palliser, qui est sensible aux ~Ccheresses,les eaux souterraines constituent
une resource importante mais ma1 dClirnit6e. La population non urbaine de la rCgion dCpend en bonne
partie des eaux souterraines pour son approvisionnement en eau. L'Ccoulement des eaux souterraines
fournit un flux de base A de nombreuses masses d'eau de surface, y compris aux lacs peu profonds. Le substratum rocheux et les dCp6ts glaciaires constituent d'importants aquifkres, la qualit6 de l'eau et la
sensibilitC aux fluctuations climatiques, A la surexploitation et h la contamination Ctant variables.
Ironiquement, ce sont les aquifkres les plus vulnCrables qui contiennent l'eau de meilleure qualitC. Les eaux
souterraines reprCsentent une source potentielle d'archives sur les fluctuations du climat dans le passC, cela
directement, sous forme de concentrations isotopiques stables, et indirectement, eu Cgard-aux profondeurs
atteintes par les nappes phrCatiques en rCaction aux changements climatiques holocknes. A ce jour, aucune
observation sur de telles Cvolutions n'a CtC rCalisCe B des Cchelles temporelles qui soient utiles A la gestion
des ~Ccheresses.Les Cchelles spatiales et temporelles des systkmes d'eaux souterraines s'Ccoulant dans les
lacs ainsi que I'hydrochimie, le volume et les sources d'alimentation des eaux sont tous susceptibles
d'amortir les liens entre les indicateurs indirects des niveaux lacustres et la prCcipitation.

' Department of Geological Sciences and Geological Engineering, Queen's University, Kingston, Ontario K7L 3N6
Department of Earth Sciences, University of Waterloo, Waterloo, Ontario N2L 3G1

GSC Bulletin 534

INTRODUCTION
In this paper, we introduce the basic concepts of
hydrogeology, including the differences between the water
table and potentiometric surface, and the characteristics of
aquifers and aquitards. We seek to differentiate aquifers on
the basis of their vulnerability to contamination, overpumping, and climate change, which is a function of the
thickness and nature of the overlying material. In the Palliser
Triangle, potable water occurs in both bedrock and glacial
aquifers, although much of that water fails to meet the drinking water objectives for some natural constituents set by
Health and Welfare Canada (1993). Despite the importance
of groundwater and the considerable efforts of provincial
research agencies, the boundaries and potential yields of most
aquifers are poorly known (Pupp et al., 1989,1991; Betcheret
al., 1995). Finally, we discuss ways in which groundwater
can both help and hinder the quest for paleoclimate information in the Palliser Triangle.

DEFINITIONS
The flow of groundwater, described by Darcy's Law, is the
sum of the hydraulic gradient 'i' (m-m-l) and hydraulic conductivity 'K' (mvs-') (Freeze and Cherry, 1979). The hydraulic gradient is the driving force for flow and is the difference
in hydraulic head (the sum of pressure head and elevation
head) between two measuring points. Hydraulic conductivity
is a measure of how freely a geological body transmits water.
It varies over at least 13orders of magnitude, from highly conductive (~=lO~m.s-l)
to highly nonconductive (K<10-l3mes-l),
and is a critical but difficult measurement to make. Several
current groundwater textbooks (e.g. Freeze and Cherry,
1979; Fetter, 1994) provide tables of estimates for K in different geological materials. These values are useful only as
initial approximations, because K varies from deposit to
deposit and usually varies within a deposit. Obtaining

meaningful measurements can be problematic, since K can


change considerably with the scale at which the measurement
is made.
The water table is defined as the surface of continuous
water-filled pores where the pressure head is exactly equal to
atmospheric pressure (Fig. 1). It occurs between ground surface and the zone of groundwater, and is usually a subtle
reflection of topography (Toth, 1963). Cherry (1987) estimated that the water table is within 20 m of ground surface in
most parts of Canada. The water table is distinguished from
the potentiometric surface, which is more correctly called a
map of hydraulic heads (Freeze and Cherry, 1979) and is a
reflection of the hydraulic head within a unit (Fig. 1). Only in
a surficial deposit is the water table equal to the hydraulic
head.
Generally, deposits are classified hydrogeologically as
either aquifers or aquitards (Fig. 2). Aquifers are defined as
geological bodies that have K values sufficiently large to
allow them to transmit water freely. Conversely, aquitards
have much lower K values and do not transmit water freely. In
general, what is not an aquifer is considered to be an aquitard.
The definition of aquifer gives no indication of size or of the
volume of water available for extraction. A 'useful' aquifer is
therefore one that can supply sufficient water for a single
home, a village, or a city, depending on the identified need.
These definitions are necessarily imprecise because, if
water-supplying units are scarce in a region, what might be
used as an aquifer there may be considered an aquitard
elsewhere.

AQUIFERS AND AQUITARDS


There are two types of aquifers, unconfined and confined
(Fig. 2). Unconfined aquifers, also called water-table aquifers, occur at ground surface and are recharged directly by
meteoric precipitation. Because they are unprotected,

Figure I .
Schematic diagram of the water table in a
surficial aquifer, a confining aquitard, and
the potentiometric sugace in a confined aquifer. The inverted triangles represent the water
table or indicate a water level.

V.H. Remenda and S.J. Birks

unconfined aquifers are vulnerable to exploitation by


overpumping, changing water volumes owing to seasonal
precipitation fluctuations, and contaminant spills and leaks.

centimetres per year. More importantly there is a potential for


very rapid movement of water, and thus of contaminants,
through fractures.

Meyboom (1967) estimated that only 1% of the surficial


deposits in the glaciated plains of western Canada are
glaciofluvial deposits. It is nbt surprising, then, that confined
aquifers are much more common than unconfined aquifers.
Confined aquifers, by definition, are not directly recharged
by precipitation but are reliant on leakage from overlying
aquitards. Although these overlying units afford some measure of protection, the effectiveness of that protection
depends on aquitard K and thickness. Some bedrock aquifers
are both confined and unconfined, and thus are recharged
both directly, where they outcrop, and through leakage.

All surficial aquitards have shallow fractures that extend


to depths ranging from 5 to 15 m. Except where their surfaces
are coated with secondary minerals, visual detection of fractures is difficult. The zone in which coatings typically occur is
restricted to the top 15 m of a deposit. Fractures that extend to
depths greater than 15 m cannot be detected visually, and
their presence can only be determined indirectly by detailed
hydraulic measurement of Kb (e.g. Keller et al., 1986). A thin
aquitard, or one where there are deep, conductive fractures,
does not provide adequate protection from fluctuations in
amount of precipitation or from contamination. This is especially problematic because, until recently, most till and clay
have been thought to be impermeable.

The vulnerability of a confined aquifer to overpumping,


fluctuations in meteoric precipitation, and contamination
depends on the thicknessmd the bulk K of the overlying
aquitard. Aquitards in the Palliser Triangle are usually clayey
till or, less frequently, glaciolacustrine deposits. These
deposits may vary in thickness from a few metres to tens of
metres and, based on the experience gained at aquitard field
sites just outside the Palliser Triangle, they may exhibit K values ranging from 1x10-l1rn.s-l to 1x10-9 m-s-l.If measurements of field-scale K (Kb) are similar to measurements of
small-scale K (h)
and are uniformly low, in the range of
lo-" to 10-lo m.s-' (e.g. Remenda et al., 1996), then the
aquitard is intact and provides effective protection. More
commonly, however, deep fractures augment Kb, causing it
to be as much as
m s l . Although this increase in K may
not appear to be significant, it translates into bulk rates of
groundwater movement of metres per year instead of

Despite the fact that there are only a small number of well
characterized field sites on the Prairies (Keller et a]., 1986;
Hendry et al., 1986; Remenda, 1993), the results of these
studies indicate that the occurrence of deer, fractures is the
norm rather than the exception. It should be noted that there
are field sites where the absence of deep fractures has been
demonstrated (Remenda et al., 1994, 1996). Although
researchers in aquitard hydrogeology continue to speculate
about the origin of deep fractures, shallow fractures (el5 m)
are generally attributed to desiccation. However, there are
currently no hypotheses that provide a satisfactory explanation for all field observations, many of which are conflicting.
For example, while McKay and Fredericia (1995) attributed
the origin of shallow fractures observed at field sites in southwestern Ontario to desiccation, a more detailed investigation
of the same area by Klint (1996) described several sets of

Figure 2.
Representation of unconfined and confined
aqugers in glacial and bedrock deposits.
Aquitards are the units composed of till, clay,
and shale.

g$,,$#

stxi;
1
3
,p,,

$,%..

..,

Sand & gravel


Shale

>,,,,

- -Clay

Till
Sandstone

GSC Bulletin 534

fractures, not all of which appeared to result from desiccation.


It is the opinion of the authors that deep fractures likely relate
to the extrusion of glacial meltwater from deposits underlying
the glacier sole due to pressure exerted by the overlying ice.
This hypothesis, proposed by Boulton and Caban (1995), is
controversial and has yet to be tested.

GROUNDWATER AS A RESOURCE
Groundwater is an extremely important resource in the
Palliser Triangle. In both Saskatchewan and Manitoba, 90%
of rural water supply is groundwater (Hess, 1981). About
25% of the populations of Alberta and Manitoba, and 50% of
the population of Saskatchewan, rely on groundwater (Hess,
1981). In the Palliser Triangle, where surface water is scarce
and precipitation is lowest, it is reasonable to assume that the
percentage of population reliant on groundwater is as great as,
or greater than, the figures given above for the provinces as a
whole.
Less obvious, but no less important, is that groundwater
supplies base flow to surface water. Lennox et al. (1988)
noted that buried-valley aquifers often drain to present-day
rivers. Changes in groundwater levels may have an impact on
surface-water systems that range in size from major rivers
(e.g. South Saskatchewan River) to small-scale wetlands. A
long period of decreased precipitation will result in an associated but not necessarily synchronous decrease in surface-water levels. The magnitude of that impact depends not
only on the size of the surface-water body, but also on the type
of groundwater system (shallow, intermediate, or deep) feeding it. During the drought of the 1980s, although water tables
declined, the water levels in deep aquifers monitored by the
SaskatchewanResearch Council continued to rise (G,van der
Kamp, pers. comm., 1998). Thus, base flow from such deep
aquifers was unaffected.
Unlike other extractable subsurface resources such as
petroleum and minerals, groundwater is renewable and,
according to Karvinen and McAllister (1994), is considered a
free good. In reality, however, groundwater is neither perfectly renewable nor absolutely free. Groundwater may be
irreparably contaminated, may be removed in excessive
quantities so that there is a decrease in the actual pore space in
the aquifer, or may be removed altogether when surficial
aquifers are mined for aggregate. There are costs associated
with groundwater delineation, extraction (both equipment
and energy), treatment, contaminant assessment and cleanup,
and delivery infrastructure. But, in Canada, as far as we are
aware, there are no charges levied for the water itself. In the
United States, there is currently a move to place an economic
value on groundwater, which has traditionally been considered both free and invaluable (National Research Council,
1997). On the one hand, groundwater has been assigned a
small or negligible value (i.e. cost per volume used) as an
extractable resource, while on the other hand, governments
have invested billions of dollars in efforts to clean up contarnination (National Research Council, 1997). This conundrum
encapsulates the notion that access to clean, low- or no-cost
water is an inherent right.

In Canada, there are both federal and provincial regulations that govern groundwater. However, the primary responsibility for sustaining quality and quantity, for maintaining
groundwater databases, for permitting extraction and
well-drilling, and for pollution prevention lies with the provincial governments. It is interesting to note that, in both
Saskatchewan and Alberta, where surface water is scarce,
provincial agencies have been active in assessment of
groundwater resources since the 1950s. In Manitoba, similar
efforts began in the 1960s. Groundwater protection legislation, as noted in the excellent review of groundwater policy
by Karvinen and McAllister (1994), is much more recent.
While all of the major aquifers in the Palliser Triangle
have been mapped and tested, many others have not been
mapped in detail, probably because aquifer delineation is an
expensive and time-consuming process. All three provinces
have extensive databases of testhole and water-well logs, and
government agencies continue to provide excellent maps of
groundwater resources. The reviews of groundwater quality
and management for Alberta (Pupp et al., 1989), Saskatchewan (Pupp et al., 1991), and Manitoba (Betcher et al., 1995),
initiated by the National Hydrology Research Institute, provide names of provincial agencies and contacts.
The chemistry, and thus the natural quality, of groundwater depends on the supply of reactants and the time since
recharge. In unconfined aquifers, water is generally young
and fresh, whereas in confined aquifers, water is older and,
because it has been in contact with till, brackish. Total dissolved solids (TDS) is an indicator of the water quality. Many
groundwater supplies in the Palliser Triangle do not meet the
water quality objective (TDSs5OO m g . ~ - l set
) by Health and
Welfare Canada (1993). Indeed, much of the groundwater in
use fails to meet the drinking water objectives for naturally
occurring constituents such as sulphate or calcium set by
Health and Welfare Canada (1993). Low-TDS water is
important for drinking water, irrigation, and industrial
supply.

Bedrock aquifers
In the Palliser Triangle, important bedrock aquifers occur
within Cretaceous and Tertiary deposits (Fig. 3). Because of
their depth and residence times, bedrock aquifers are the least
vulnerable to contamination, overpumping, and climate
change. The degree to which these aquifers have been investigated varies from intensively, like the Milk River Aquifer in
southern Alberta, to not at all, like the locally important but
unmapped Tertiary aquifers in the Cypress Hills (Pupp et al.,
1991). It is not surprising, then, that information regarding
well yields (i.e. the amount of water that can be pumped from
the aquifer) and water quality is scarce. In some cases, the
aquifers have been named only informally, and different
names may be used locally. In southern Alberta, the most
important aquifer is the Milk River Aquifer, which is a marine
sandstone that was originally mapped and evaluated by
Meyboom (1960). The water quality in the Milk River Aquifer
ranges from 1000 to 2500 m g . ~ - TDS.
l
Pupp et al. (1989)
noted several nonmarine sandstone members that provide
locally important aquifers. For example, in the Medicine Hat

V.H. Remenda and S.J. Birks

area, most of the water-producing wells are completed in the


Belly River Group, where TDS ranges from 1000 to
3000 r n g - ~ -(Stevenson
'
and Borneuf, 1977).
Pupp et al. (1991) identified the Judith River Formation
(equivalent to the Belly River Group in Alberta) as the lowermost important aquifer in Saskatchewan. Overlying the
Judith River Formation is the Bearpaw Formation, a thick
regional aquitard composed primarily of poorly indurated silt
and clay. Although sand members of the Bearpaw Formation
form locally important aquifers in Saskatchewan, they have
not been extensively mapped. The Ravenscrag Aquifer, comprising beds from both the Upper Cretaceous Eastend and
Tertiary Ravenscrag Formations, outcrops in southwestern
Saskatchewan and subcrops below glacial deposits in southeastern Saskatchewan (Pupp et al., 1991).
In Saskatchewan, TDS reported for water from the Judith
River Formation ranges from 1300 to 8000 m g . ~ - l(Pupp
et al., 1991). Where the aquifer is used extensively, however,
the water quality is generally good, with TDS of 2500 mg.L-'
or less (Pupp et al., 1991). The water quality in the
Ravenscrag Aquifer ranges from 1000 mg.L-I TDS in the
western area where it outcrops, to 2600 mg-L-' TDS in the
eastern area where it is overlain by till (Pupp et al., 1991).
In southwestern Manitoba, the most important bedrock
aquifer is found in the fractured Odanah Member of the Pierre
Formation, which is commonly called the Pierre Shale.
Beyond the easternmost extent of the Judith River Formation,
the Cretaceous shales are called the Pierre Formation.
Domestic wells completed in the Odanah Shale commonly

Period

Southern
Alberta
Drift

a,

Southern
Saskatchewan
Drift

Southwestern
Manitoba

Drift

Wood Mountain Fm.1

.-a

Cypress
Hills Fm.

River Fm.

Cypress
Hills Fm.
Swift Current Fm
I ~ a v e n s c-r aFm
~
.-I

Bearpaw Fm.

piiiiK4

V)

=I

a,

&
%
a-

a s?

30

Pierre Fm.
PakowkiFm.

PakowkiFm.

Figure 3. Generalized bedrock stratigraphy of the Palliser


Triangle. Boxed units are aquifers.

yield less than 0.5 L-s-I, but in some cases may yield 1 L.s-l.
The water quality is variable, with TDS ranging from
500 m g - ~ -tol 9000 m g . ~ -TDS
'
(Betcher et al., 1995).

Non-indurated deposits
There are three types of aquifers associated with late
Tertiary-early Quaternary deposits: buried-valley aquifers,
intertill aquifers, and surficial aquifers. Buried-valley and
intertill aquifers are confined by lower K till, whereas
surficial aquifers occur at ground surface (Fig. 2).
Buried-valley aquifers are usually regionally extensive and
are composed of sand and gravel belonging to the Saskatchewan Group in Alberta and the Empress Group in Saskatchewan (Whitaker and Christiansen, 1972). Intertill aquifers
vary considerably in size, and can be both local and regional
in extent. Surficial aquifers are generally only locally important. Lennox et al. (1988) noted that there can be complex
hydraulic connections between intertill and buried-valley
aquifers, and also between buried-valley and bedrock aquifers.
Natural water quality is variable, but is generally very
good in surficial aquifers and poorer in buried-valley and
intertill aquifers. Freeze and Cherry (1979) described two
water types, Type I1 and Type 111, that occur in glacial deposits on the Prairies. Type I1 waters are Ca-Mg-HC03 type, with
TDS less than 1000 mg-L-l.Type I1 waters are found in local,
surficial aquifers and, because they have little contact with
till, they provide the best quality water. Type 111 waters are
Na-Ca-Mg-SO4 type, with TDS ranging from 1000 to
10 000 mg.L-l. Type I11 waters are found in intertill and
buried-valley aquifers. It is contact with till, and the
hydrochemical reactions occurring within the till (including
dissolution of carbonates, oxidation of sulphides, and cation
exchange (Keller and van der Kamp, 1988)), that cause the
high TDS in Type I11 waters. Fortin et al. (1991), working on
an extensive intertill aquifer in the Saskatoon area, observed
that the aquifer was simply a mixing zone for water that had
developed its chemistry while moving through the overlying
aquitard, and that high TDS is correlated with high SO4, Ca,
and Mg. Because the contacting till has a nearly inexhaustible
supply of calcium-, magnesium-, and sulphur-bearing rninerals, most intertill and buried-valley aquifers have excessive
concentrations of calcium and magnesium (hardness), and of
sulphate.
Betcher et al. (1995) and Pupp et al. (1989, 1991) indicated that buried-valley aquifers are an important groundwater resource. A typical, and regionally important, example is
the Estevan Valley Aquifer near the city of Estevan in southeastern Saskatchewan. It is a long, thin, sinuous 'pipe' of sand
and gravel of the Empress Group (Whitaker and Christiansen,
1972), overlain by thick till. Similar buried-valley aquifers
are known to occur in southwestern Manitoba and southeastern Alberta, although their boundaries and yields are imperfectly known (Pupp et al., 1989; Betcher et al., 1995).
Because they are overlain by thick till aquitards, recharge to
buried-valley aquifers is very slow and buffered from even
decade-scale changes in precipitation.

GSC Bulletin 534

Intertill aquifers vary widely in frequency, thickness, and


areal extent. Although there are numerous locally important
intertill aquifers in the Palliser Triangle, many of these have
not been mapped in detail (Pupp et al., 1989, 1991; Betcher
et al., 1995). Depending on the thickness and nature of the
overlying aquitard, intertill aquifers may or may not be vulnerable to climate change.
Surficial aquifers are more likely to be locally, rather than
regionally, important and not all have been mapped (Pupp
et al., 1989, 1991; Betcher et al., 1995). Without a lower permeability confining layer, they yield high-quality water, but
are also vulnerable to overpumping, drought, and contamination.

GROUNDWATER AND
PALEOCLIMATE STUDIES
In the drought-prone Palliser Triangle, laowledge of the temporal scales of climate fluctuations and the associated
response of the region is important. There are two ways in
which groundwater may be important in paleoclimate studies. First, groundwater may provide information about
changes in temperature and amount of precipitation. Second,
groundwater may interfere with the connection between
paleolimnological proxy indicators and paleoclimate.

Groundwater as an archive of paleoclimate


information
Groundwater or groundwater conditions can provide an
archive of paleoclimate information in two ways. First, information about air temperature during recharge can be inferred
from stable isotope concentrations along flow paths within a
groundwater system. Second, by comparing present-day
water tables to evidence of paleo-water tables, qualitative
information is obtained about the decrease in precipitation
volumes.
Paleotem~eratureinformation can be inferred from the
ratio of the stable isotopes of hydrogen and oxygen (Z2H,
6180; expressed in the per mil notation), which are present in
all water, including groundwater. Values for F2H and 6180 in
precipitation sampled throughout the world are empirically
correlated by the equation for the modern meteoric water line
(F2H = 88180 + 10; Craig, 1961). On the global meteoric
water line, more negative values are associated with colder
temperatures, and more positive values with warmer temperatures. On a local meteoric water line for a given area, more
negative values are associated with winter precipitation, and
more positive values with summer precipitation.
In using stable isotopes in groundwater to investigate
paleoclimate, one is loolcing for more positive (implying
warmer climate conditions) or more negative (implying
colder climate conditions) values, preserved along the flow
path, that can be correlated temporally with some fluctuation
in climate. Although both aquifers (Bath, 1983; Rozanski,
1985; Phillips el al., 1986; Stute et al., 1992; Stute and
Schlosser, 1993; Stute et al., 1995) and aquitards (Remenda
et al., 1994) have the potential to preserve a record of

precipitation infiltrated at the recharge area, only dramatic


climate changes such as the large temperature shift during the
Pleistocene-Holocene transition have thus far been identified. Fluctuations on a finer temporal scale have not yet been
observed, although Birks et al. (in press) predicted, on the
basis of groundwater modelling, that 6 2 and
~ 6180 contents
sampled from shallow depths in aquitards have potential to
preserve smaller scale signals. A follow-up field study is
underway.
Whether the surficial deposit is aquitard or aquifer, the
water table rises and falls in response to seasonal and longer
term fluctuations in precipitation and evapotranspiration. It
seems likely that the magnitude of the change (i.e. how far the
water table drops) is also related to K and landscape position.
Both stratigraphic and hydrogeological studies of tills commonly report the depth from ground surface to the interface
between the weathered zone and the unweathered zone ( e . ~ .
Christiansen, 1992; Remenda, 1993), which, for the purpbsG
of this paper, we will call the depth of the oxidized zone. The
oxidized zone is separated from the unoxidized zone on the
basis of colour, determined by comparing wet till samples to a
MunsellTMColour Chart. For four field sites situated in till in
Saskatchewan, samples taken from oxidized zones are usually 2.5Y 512 (greyish brown) or 5Y 512 (olive grey), whereas
samples from unoxidized zones are usually 5Y 511 (grey)
(Remenda, 1993). Christiansen (1992) gave a comprehensive
review of colours and descriptions of oxidized zones
Generally, the depth of the oxidized zone exceeds the
depth of the current water table and has been attributed to loss
of moisture and influx of oxygen during a period of warmer,
drier conditions, presumably during the mid-Holocene (ca.
7 5 0 0 4 0 0 0 BP). At field sites in southern Saskatchewan, oxidized zones up to 13 m deep have been observed in aquitards
(Keller et al., 1986; Remenda, 1993). In similar deposits in
southwestern Manitoba, oxidized zones occur to an average
depth of 6 m (R.N. Betcher, pers. comm., 1998). At several
aquitard field sites in southeastern Alberta, the depth of the
oxidized zone exceeds 15 m (Hendry et al., 1986). It should
be noted that, although this is not a comprehensive survey of
field sites throughout the Palliser Triangle, it does appear that
there is an increase in the depth of the oxidized zone from 6 m
in the east to at least 15 m in the west.
If, at a given aquitard site, the depth of the oxidized zone is
assumed to correlate with the depth to which the water table
declined during the mid-Holocene, and is not influenced by
weathering from a previous interglacial, then the decline in
regional water tables may have been 6-15 m. This is consistent with, but better constrained than, the estimate of a decline
of greater than 4 m made by Lemmen et al. (1997) on the basis
of lake sediment investigations.

Groundwater interference with paleoclimate proxy


indicators
Paleoclimate information may be preserved in the sediments
of closed-basin, groundwater-fed lakes. The assumption is
that the occurrence of certain mineral assemblages and
fossils, and the thickness of sediment layers, are indicators of

V.H. Remenda and S.J. Birks

past lake levels. Thus, changes in the proxy indicators are


equated with past changes in lake levels, generally interpreted
to reflect changes in precipitation and evaporation. This is
predicated on the assumption that precipitation and evaporation are the main controls on lake levels and lake chemistry,
and that groundwater inputs are either constant or synchronous with precipitation fluctuations (Vance et al., 1992). If,
however, the volume of groundwater and the source of
groundwater have not been constant or synchronous with climate change, this assumption is not valid.
The influence of groundwater on lakes is reflected by the
type of groundwater system discharging to the lake (i.e. some
combination of local, intermediate, or deep systems that varies in time), the volume of water gained or lost, andlor the
chemical mass gained or lost. Lakes most sensitive to climate
are steep-sided, closed basins in arid to subhumid areas where
g-oundkater transfer is negligible (Street-Perrott and Harrison, 1985). In the Palliser Triangle, however, most lakes are
shallow, evaporative, and fed by groundwater (Last, 1999).
These lakes may be fed by local (surficial aquifers, shallow
fractured aquitards), intermediate (intertill aquifers, shallow
buried-valley aquifers), or deep (deep buried-valley aquifers,
bedrock aquifers) groundwater systems, depending on their
position in the flow regime. Local, intermediate, and deep
systems have different residence times and may have different chemical compositions (Winter, 1976; Wallick, 1981;
Donovan and Rose, 1994). It is impossible to determine with
confidence the type of system that is feeding a lake without a
groundwater investigation (e.g. Birks and Remenda, 1999).
Toth (1963) developed models of these three systems on
the basis of the steady-state analytical solutions to the
groundwater-flow equation using two-dimensional, homogeneous basins with varying water table and impermeable
boundaries. These simple scenarios are useful for making
generalizations about flow systems, although they do not
account for complex stratigraphy.
With short residence times, high flow rates, and an effective base of influence from 30 to 45 m (Toth, 1962, 1963),
local groundwater systems have the most direct but not necessarily synchronous connection with climate fluctuations. The
delay between change in precipitation volume and change in
the system response depends on K, where the lower the K
value, the larger the lag time. FIow is driven by highs and
lows in topography, and flow vectors intercept land surface
and the water table obliquely or almost horizontally (Toth,
1963). Weathering, hydrolysis reactions, and the dissolution
of easily soluble minerals (carbonates and sulphates) are
dominant in these systems (Donovan and Rose, 1994).
An intermediate groundwater system spans two or more
local systems. The driving force for flow is the difference in
fluid potential between
highs and lows owing to
incised
Or
(Donovan
and Rose, 1994). The intermediate system is farther removed
surface processes and may not reflect current 'limate
fluctuations. Maathuis and van der Kamp (1986), in their
analysis of data from observation wells in intertill and buried-valley aquifers, noted that the potentiometric surface, and
thus the discharge to surface water, are apparently unaffected

by climate fluctuations where the overlying aquitard is thick


and unfractured. There are insufficient data (only 30 years) to
determine whether this holds true over time periods longer
than decades. The water quality in intermediate systems is
similar to that of shallow systems, dominated by carbonates
and sulphates, but the TDS is higher.
Deep groundwater systems have the longest residence
times and lowest flow rates, and are therefore unlikely to
reflect current climate fluctuations. With bedrock aquifers,
discharge occurs only at the edge of basins at major groundwater divides. Although discharge areas are often shared by
local, inter~nediate,and deep systems, flow vectors from deep
systems are vertical or near vertical (Toth, 1963). These long
residence times usually result in a more concentrated chemical composition that trends toward sodium chloride or
sodium sulvhate waters.
Winter (1976) analyzed the flow of groundwater in the
vicinity of a lake, and determined that it is sensitive to variations in hydrogeological setting. Using two-dimensional,
steady-state numerical simulations, he found that the geometry and K of the geological medium, height of the water table,
and depth of the lake strongly influence the position, shape,
and continuity of the flow field. Slight variations in
hydrogeological setting can lead to different interactions
between groundwater and surface water. For example, a local
flow system can be entirely separate from the regional flow,
so that there is no transfer of water from one system to the
other (Winter, 1976). Shallow groundwater enters the lake at
the edge and not through the base. In contrast, if the local and
regional systems are not isolated, then the lake loses or gains
water, usually through its base. Loss to underlying aquifers,
in a field situation similar to those approximated by the Winter model, has been reported by Shaw and Prepas (1990).
The hydrochemistry of prairie lakes ranges in concentration and composition from dilute, Ca-HC03 water to
hypersaline, Na-SO4 brine (Teller and Last, 1990; Last, 1992;
Donovan and Rose, 1994; Last, 1999). The evolution of
lake-water chemist~yis a function of evaporative concentration, which in turn depends on the initial composition, climate
(evaporation versus precipitation), and lake hydrology
(groundwater-surface-water interactions) (Wood and
Sanford, 1990; Last, 1992; Donovan and Rose, 1994). Several researchers have speculated on the source of solutes for
prairie lakes and hypotheses include: dissolution of deep
Paleozoic evaporites by upward-flowing water (Grossman,
1968); evaporation of groundwater from shallow Cretaceous
and Tertiary bedrock sources (Wallick, 1981; Last, 1992);
and evaporation of groundwater from shallow surficial aquifers (witkind, 1952; Donovan and Rose, 1994). On the basis
of recent research in till hydrogeology and hydrochemistry
(Keller and van der Kamp, 1988; Fortin et al., 1991), it is
likely that flow of highly mineralized groundwater from the
weathered zone of ti]] units
significantly to the
chemical mass in lakes (see
Last, 1999). The contribution from weathered till has probably been underestimated.
The advection of lake water from below the lake
described by Winter (1976) is one mechanism by which
seemingly closed lakes can lose water and solute. Molecular

GSC Bulletin 534

diffusion and density-driven flow are two other means by


which closed-basin lakes can lose solute (Wood and Sanford,
1990). If the lake water becomes a brine, then the equivalent
fresh-water hydraulic head within the lake exceeds the
fresh-water hydraulic heads in the basin below and fingers of
brine can sink into the underlying units (Wood and Sanford,
1990; Ferguson et al., 1992; Birks and Remenda, 1999).
Although the vertical hydraulic conductivity of the lake floor
sediments controls this process, density differences and evaporation rates are also important (Ferguson et al., 1992). For
saline lakes with low-permeability deposits at their base, diffusion of solute into surrounding media with lower solute
concentrations can cause significant solute loss (Sanford et
al., 1989; Ferguson et al., 1992). Although groundwater can
contain information about paleoclimate, it is more likely in
the Palliser Triangle that groundwater can interfere with the
paleoclimate record contained in lake sediments. Many lakes
are not strictly 'closed basin', as they may lose or gain water
and chemical mass owing to interactions with groundwater
(e.g. Birks and Remenda, 1999). These interactions are site
specific.

SUMMARY AND CONCLUSIONS


Groundwater is an important resource in the Palliser Triangle. A large percentage of the population relies on groundwater for domestic purposes, municipal supplies, stock
watering, and irrigation. The best quality groundwater is
found in surficial sand and gravel, and is contained in the
aquifers most vulnerable to climate change, overpumping,
and contamination. Poorer quality water in intertill and buried-valley aquifers is more abundant and, depending on the
thickness and bulk K of the overlying aquitard, less vulnerable to drought and contamination. The origins of fractures
that increase the bulk K of aquitards are still unclear. Despite
its importance and the efforts of provincial bodies, knowledge of the nature and extent of accessible groundwater is still
incomplete.
Groundwater conditions may provide climate information. Paleotemperature information can be inferred from
changes in the 6 2 and
~ 6 ' 8 0 content of groundwater along a
flow path. Qualitative information may be obtained from a
survey of the depth from ground surface to the interface
between the zones of oxidized and unoxidized till. Although
not a comprehensive survey, it appears that there is an
increase in the depth of the oxidized zone from 6 m in the east
to at least 15 m in the west, which may correlate with precipitation trends. The temporal scale may, however, be too coarse
to be of use in planning strategies to combat shortages.
Groundwater also provides base flow to surface-water
courses, and groundwater-fed lakes are common in the
Palliser Triangle. The spatial and temporal scales of the
groundwater systems, the hydrochemistry, the volume of
water, and the source of water all have the potential to 'nullify' the connection between proxy indicators of lake level
and precipitation. Thus, climate proxies from lake sediments
must be carefully examined to ensure that groundwater does
not obscure them.

ACKNOWLEDGMENTS
The authors thank Bill Last and Bob Vance for their constructive and thoughtful reviews. The considerable efforts of
Don Lemmen and Bob Vance for their time and patience in
bringing this volume to fruition is gratefully acknowledged.

REFERENCES
Bath A.H.
1983: Stable isotopic evidence for palaeo-recharge conditions of groundwater; in Palaeoclimate and Palaeowater: A Collection of Environmental Isotope Studies; International Atomic Energy Agency,
Vienna, Austria, p. 169-186.
Betcher, R., Grove, G., and Pupp, C.
1995: Groundwater in Manitoba: hydrogeology, quality concerns, management; Environment Canada, National Hydrology Research
Institute, Contribution No. CS-93017,47 p.
Birks, S.J. and Remenda, V.H.
1999: Groundwater inputs to a closed-basin saline lake, Chappice Lake,
Alberta: in Holocene Climate and Environmental Change in the
~ a l l i s e Triangle:
i
A Geoscientific Context for Evaluating the
Imoacts of Climate Change on the Southern Canadian Prairies. led.)
D.S. Lemmen and ~.E:Vance; Geological Survey of canada;
Bulletin 534.
Birks, S.J., Edwards, T.W.D., and Remenda, V.H.
i n press: Clay aquitards as isotopic archives of Holocene paleoclimate in the
northern Great Plains; Journal of Hydrological Processes.
Boulton, G.S. and Caban, P.
1995: Groundwater flow beneath ice sheets: part II - its impact on glacier
tectonic structures and moraine formation; Quaternary Science
Reviews, v. 14, p. 563-587.
Cherry, J.A.
1987: Groundwateroccurrence and contaminationin Canada; inCanadian
Aquatic Resources, (ed.) M.C. Healey and R.R. Wallace; Rawson
Academy of Aquatic Science, Canadian Bulletin of Fisheries and
Aquatic Sciences, v. 215, p. 387-426.
Christiansen, E.A.
1992: Pleistocene stratigraphyof the Saskatoon area, Saskatchewan,Canada: an update; Canadian Journal of Earth Sciences, v. 29,
p. 1767-1778.
Craig, H.
1961: Isotopic variations in meteoric waters; Science, v. 133,
p. 1702-1703.
Donovan, J.J. and Rose, A.W.
1994: Geochemical evolution of lacustrine brines from variable-scale
groundwater circulation; Journal of Hydrology, v. 154, p. 35-62.
Ferguson, J., Jacobson, G., Evans, W.R., White, I., Wooding, A.,
Barnes, C.J., and Tyler, S.
1992: Advection and diffusion of groundwater brines in modem and
ancient salt lakes, Nulla groundwater discharge complex, Murray
Basin, southeast Australia; in Proceedings, 7Ih International Symposium on Water-Rock Interaction, Park City, Utah, July 1992,
(ed.) Y.K. Kharaka; A.A. Balkema, Rotterdam, The Netherlands
and Brookfield, Vermont, p. 643-647.
Fetter, C.W.
1994: Applied Hydrogeology; Maxwell Macmillan Canada, Toronto,
Ontario, 691 p. (third edition).
Fortin, G., van der Kamp, G., and Cherry, J.A.
1991: Hydrogeology and hydrochemistry of an aquifer-aquitard system
within glacial deposits, Saskatchewan, Canada; Journal of Hydrology, v. 126, p. 265-292.
Freeze, R.A. and Cherry, J.A.
1979: Groundwater;Prentice Hall, Englewood Cliffs, New Jersey, 604p.
Grossman, I.G.
1968: Origin of the sodiumsulfate deposits of the northern Great Plains of
Canada and the United States; United States Geological Survey,
Professional Paper 600-B, p. B104-B109.

V.H. Remenda and S.J. Birks

Health and Welfare Canada


1993: Guidelines for Canadian drinking water quality; prepared by the
Federal-Provincial Subcommittee on Drinking Water of the Federal-Provincial Advisory Committee in Envkonmental and Occupational Health, Supply and Services Canada, Ottawa, 24 p. (fifth
edition).
Hendry, M.J., Cherry, J.A., and Wallick, E.I.
1986: Origin and distribution of sulfate in a fractured till in southern
Alberta, Canada; Water Resources Research, v. 22, p. 45-61.
Hess, P.J.
1981: Ground-water use in Canada; Environment Canada, Inland Waters
Directorate, Technical Bulletin 140,43 p.
Karvinen, W.O. and McAllister, M.L.
1994: Rising to the surface: emerging groundwater policy trends in Canada; Centre for Resource Studies, Queen's University, Kingston,
Ontario, 143 p.
Keller, C.K. and van der Kamp, G .
1988: Hydrogeology of two Saskatchewan tills, LI. Occurrence of sulfate
and implication for soil salinity; Journal of Hydrology, v. 101,
p. 123-144.
Keller, C.K., van der Kamp, G., and Cherry, J.A.
1986: Fracture permeability and groundwater flow in clayey till near
Saskatoon, Saskatchewan; Canadian Geotechnical Journal, v. 23,
p. 229-240.
Klint, K.E.S.
1996: Fractures and depositional features of the St. Joseph Till and upper
part of the Black Shale Till at the Laidlaw site, Lambton County,
Ontario; Geological Survey of Denmark and Greenland, Report
1996/9,39 p.
Last, W.M.
1992: Chemical compositionof saline and sub-salinelakes of the northern
Great Plains, western Canada; International Journal of Salt Lake
Research, v. I, no. 2, p. 47-76.
1999: Geolimnology of the Great Plains of western Canada; in Holocene
Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian prairies, (ed.) D.s. Lemmen and
R.E. Vance: Geological Survev of Canada, Bulletin 534.
Lemmen, D.S., ~ a n c eR.K,
,
Wolfe, KA., and Last, W.M.
1997: Impacts of future climate change on the southern Canadian Prairies:
a paleoenvironmental perspective; Geoscience Canada, v. 24,
p. 121-133.
Lennox, D.H., Maathuis, H., and Pederson, D.
1988: Region 13, western glaciated plains; in The Geology of North
America, Hydrogeology; Geological Society of America, v. 0-2,
p. 115-128.
Maathuis, H. and van der Kamp, G.
1986: Observation well network in Saskatchewan; in Drought, the
Impending Crisis? Proceedings of theCanadian Hydrology Symposium No. 16, Regina, Saskatchewan, June 3-6, National Research
Council, Associate Committee on Hydrology, Ottawa, Ontario,
p. 565-581.
McKay, L.D. and Fredericia, J.
1995: Distribution, origin, and hydraulic influence of fractures in a
clay-rich glacial deposit; Canadian Geotechnical Journal, v. 32,
p. 957-975.
Meyboom, P.
1960: Geology and groundwater resources of the Milk River sandstone in
southern Alberta; Alberta Research Council, Memoir 2,89 p.
1967: Interior plains hydrogeologicalregion; in Groundwater in Canada,
(ed.) I.C. Brown; Geological Survey of Canada, Economic Geology
Report 24, p. 131-158.
National Research Council
1997: Valuing Groundwater: Economic Concepts and Approaches;
National Academy Press, Washington, D.C., 189 p.
Phillips, F.M., Peeters, L.A., Tansey, M.K., and Davis, S.N.
1986: Paleoclimatic inferences from an isotopic investigation of groundwater in the central San Juan Basin, New Mexico; Quaternary
Research, v. 26, p. 179-193.

Pupp, C., Maathius, H., and Grove, G.


1991: Groundwater quality in Saskatchewan: hydrogeology, quality concerns, management; Environment Canada, National Hydrology
Research Institute, Contribution No. CS-91028, 66 p.
Pupp, C., Stein, R., and Grove, G.
1989: Groundwater quality in Alberta: hydrogeology, quality concerns,
management; Environment Canada, National Hydrology Research
Institute, Contribution No. 8905 1, 113 p.
Remenda, V.H.
1993: Origin and migration of natural groundwater tracers in thick clay
tills of Saskatchewanand theLake Agassizclay plain; Ph.D. thesis,
University of Waterloo, Waterloo, Ontario, 270 p.
Remenda, V.H., Cherry, J.A., and Edwards, T.W.D.
1994: Isotopic composition of old ground water from Lake Agassiz:
implications for late Pleistocene climate; Science, v. 266,
p. 1975-1978.
Remenda, V.H., van der Kamp, G., and Cherry, J.A.
1996: Use of vertical profiles of 8180 to constrain estimates of hydraulic
conductivity in a thick, unfractured aquitard; Water Resources
Research, v. 32, no. 10, p. 2979-2987.
Rozanski, K.
1985: Deuterium and oxygen-18 in European groundwater - links to
atmospheric circulation in the past; Chemical Geology (Isotope
Geoscience Sections), v. 52, p. 349-363.
Sanford, W.E., Wood, W.W., and Konikow, L.F.
1989: Groundwater flow and solute Wansport at a playa lake on the southern High Plains; American Geophysical Union Transactions, v. 70,
p. 1101.
Shaw, R.D. and Prepas, E.E.
1990: Groundwater-lake interactions: 11. Nearshore seepage patterns and
the contribution of groundwater to lakes in central Alberta; Journal
of Hydrology, v. 119, p. 121-136.
Stevenson, D.R. and Borneuf, D.M.
1977: Hydrogeology of the Medicine Hat area, Alberta; Alberta Research
Council, Report 75-2, 11 p.
Street-Perrott, F.A. and Harrison, S.P.
1985: Lake levels and climate reconstruction; irz Paleoclimate Analysis
andModelling, (ed.) A.D. Hecht; John Wiley and Sons, New York,
New York, p. 291-340.
Stute, M. and Schlosser, P.
1993: Principles and applications of the noble gas paleothermometer; in
Climate Changes in Continental Isotopic Records; Geophysical
Monograph, v. 78, p. 89-100.
Stute, M., Clark, J.F., Schlosser, P., Broecker, W.S., and Bonani, G.
1995: A 30,000 year continental paleotemperature derived from noble
gases dissolved in groundwaterfrom the San Juan Basin, New Mexico; Quaternary Research, v. 43, p. 209-220.
Stute, M., Schlosser, P., Clark, J.F., Broecker, W.S.
1992: Paleotemperatures in the southwestern United States derived from
noble gases in groundwater; Science, v. 256, p. 1000-1003.
Teller, J.T. and Last, W.M.
1990: Paleohydrologicalindicators in playas and salt lakes with examples
from Canada, Australia, and Africa; Palaeography,
Palaeoclimatology,and Palaeoecology, v. 76, p. 215-240.
Toth, J.
1962: A theory of groundwater motion in small drainage basins in central
Alberta, Canada; Journal of Geophysical Research, v. 67,
p. 43754387.
1963: A theoretical analysis of groundwater flow in small drainage basins;
Journal of Geophysical Research, v. 68. p. 47954812.
Vance, R.E., Mathewes, R.W., and Clague, J.J.
1992: A 7000 year record of lake-level change on the northern Great
Plains: a high-resolution proxy of past climate; Geology, v. 20,
p. 879-882.
Wallick, E.I.
1981: Chemical evolution of groundwater in adrainage basin of Holocene
age, east-centWal Alberta, Canada; Journal of Hydrology, v. 54,
p. 24-283.
Whitaker, S.H. and Christiansen, E.A.
1972: TheEmpress Group in southern Saskatchewan;Canadian Journal of
Earth Sciences, v. 9, p. 353-360.

GSC Bulletin 534


Winter, T.
1976: Numerical simulation analysis of the interaction of lakes and
groundwater; United States Geological Survey, Professional Paper
1001.45 p.

Witkind, I.J.
1952: Thelocalization of sodiu~nsulfate deposits in northeastern Montana
and northwestern North Dakota; American Journal of Science,
v. 250, p. 667-676.
Wood, W.W. and Sandford, W.E.
1990: Ground water control of evaporite deposition; Economic Geology,
v. 85, p. 1226-1235.

Diatom-based salinity reconstructions from


Palliser Triangle lakes: a summary of two
Saskatchewan case studies
Susan E. wilson1 and John P. smo12
Wilson, S.E. and Smol, J.P., 1999: Diatom-based salinity reconstructions from Palliser Triangle
lakes: a summary of two Saskatchewan case studies; in Holocene Climate and Environmental
Change in the Palliser Triangle: A Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological
Survey of Canada, Bulletin 534, p. 67-79.

Abstract: A diatom-based transfer function was used to infer past salinity levels for Hams Lake in the
Cypress Hills, and Clearwater Lake on The Missouri Coteau. Harris Lake has remained fresh throughout the
Holocene, with only very slight changes in salinity. A shift from a benthic to a predominantly planktonic
diatom flora between 6500 and 5200 BP may be an indirect response to a warmer climate that reduced forest
cover and allowed greater rates of inorganic sedimentation.
The early Holocene salinity reconstruction from Clearwater Lake indicates that the lake was quite saline
(4-20 g . ~ - lbetween
)
9600 and 8400 BP, similar to several other sites in the northern Great Plains, and then
became relatively fresh by ca. 7900 BP. The recent Clearwater record indicates little change in salinity over
the past 400 years. These studies also demonstrate the importance of local hydrology and anthropogenic
influences, which may at times obscure climatic signals, in regulating responses of closed-basin lakes.

RCsum6 : On a eu recours 21 une fonction de transfert basCe sur les diatomCes afin de dCduire les niveaux
de salinitC passes du lac Harris, dans les collines Cypress, et du lac Clearwater, sur le coteau du Missouri.
L'eau du lac Harris est rest6e fraiche pendant I'ensemble de 1'Holockne et n'a presentee que de tr&sfaibles
variations de salinit6. Le passage d'une flore de diatomCes benthique 9 une flore principalement
planctonique entre 6 500 et 5 200 BP est peut-&re une r6action indirecte h l'apparition d'un climat plus
chaud qui a rCduit la couverture forestikre et favoris6 des taux plus 6lev6s de sidimentation de matikre
inorganique.
La reconstitution de la salinitk du lac Clearwater 9 1'Holockne supCrieur indique que le lac Ctait
relativement salin (4-20 g . ~ - I )de 9 600 9 8 400 BP, tout cornrne de nombreux autres sites des grandes
plaines septentrionales, et qu'il s'Ctait relativement adouci vers 7 900 BP. Les donnCes rCcentes sur le lac
Clearwater indiquent que la salinitC a peu changC depuis 400 ans. Ces Ctudes montrent en outre
l'importance des influences hydrologiques et anthropogknes locales dans la rCgulation des rCactions des
lacs de bassins fermCs, influences qui peuvent parfois masquer les signaux climatiques.

' Laboratory of Paleoclimatologyand Climatology,Department of Geography,University of Ottawa, Ottawa, Ontario


K I N 6N5

Paleoecological Environmental Assessment and Research Lab (PEARL), Department of Biology, Queen's
University, Kingston, Ontario K7L 3N6

GSC Bulletin 534

INTRODUCTION
Paleolirnnology is a multidisciplinary science that uses physical, chemical, and biological information preserved in lake
sediments to reconstruct and interpret past environmental and
climatic conditions (Smol et al., 1991, 1995). The siliceous
frustules of diatoms are often the major indicators used in
paleolimnological studies, as they are abundant in most
aquatic habitats, are generally well preserved in lake sediments, and are taxonomically diagnostic (Dixit et al., 1992).
Moreover, diatoms are proven indicators of many limnological variables, such as pH (Dixit et al., 1993), total phosphorus
(Hall and Smol, 1992; Reavie et al., 1995a, b), dissolved
organic carbon (Kingston and Birks, 1990; Pienitz and Smol,
1993), temperature (Pienitz et al., 1995), and salinity (Wilson
et al., 1994; Cumrning et al., 1995).
With recent concerns about climatic change, the need for
reliable long-term paleoclimatic proxy data has been recognized by many government agencies and international organizations. Topographically closed basin, saline lakes in arid
and semiarid regions offer an excelIent setting for evaluating
fluctuations in climate, due to their strong dependence on the
balance between precipitation and evaporation
(Street-Perrott and Roberts, 1983). During intervals of
warmer and/or drier climates, lake levels generally decrease
in closed basins and brine concentrations increase (e.g.
Hammer, 1990). Conversely, during cooler and/or wetter
intervals, lake levels generally rise and salinity is reduced.
However, individual lakes can vary greatly in their sensitivity
and response to changes in moisture balance (see Fritz et al.
(1999) for a more detailed discussion). For example, lakes
having groundwater as a major component of their hydrological budget may show a limited response to climate forcing
compared to those with little groundwater inflow (Remenda
and Birks, 1999). Furthermore, nonclimatic factors, such as
infiltration of saline groundwater and threshold effects, can
also affect the salinity status of a lake and obscure the climate
signal. These problems are discussed in more detail by Gasse
et al. (1997). Changes in salinity brought about by changes in
climate or groundwater influence can have dramatic effects
on the chemistry and biota of saline lakes, and can be reconstructed using different forms of paleolimnological proxy
data, including diatoms.
The sensitivity of diatom communities to changes in
salinity and brine composition has been recognized for some
time (see review by Patrick and Reimer, 1966), and more
recent studies of athalassic (inland) saline lakes continue to
demonstrate that diatom distributions are strongly related to
lake-water ion concentration and ionic composition (e.g.
Africa (Gasse et al., 1983), the northern Great Plains of North
America (Fritz and Battarbee, 1988), the Canadian Subarctic
(Pienitz et al., 1992), Mexico (Metcalfe, 1988), South
America (Servant-Vildary and Roux, 1990), western North
America (Blinn, 1993), and Australia (Gell and Gasse,
1994)). Important insights concerning changes in Holocene
climate and salinity have been documented using qualitative
means from diatom assemblages in sediment cores from
Africa (Gasse, 1987; Gasse et al., 1987),Australia (Gell et al.,
1994), and western North America (Radle et al., 1989;

Bradbury et al., 1989; Hickrnan and Schweger, 1993). However, more objective, quantitative reconstructions of salinity
are now possible with the recent development of diatom-based transfer functions.
The basic approach used to develop a transfer function
(see review in Charles and Smol, 1994) involves choosing a
suite of study lakes (often 60-100 in number and referred to
as the training or calibration set) that span the limnological
gradients of interest (e.g. salinity, total phosphorus). Surficial
sediments representing the last few years of sediment
accumulation (e.g. the upper 0.5-2 cm, depending on the
sediment accumulation rate) are collected for diatom identification and enumeration; water chemistry and other limnological and environmental data are collected simultaneously.
Ideally, multiple samples of the limnological conditions
should be taken throughout the year to compensate for seasonal changes in water chemistry. This applies particularly to
many saline lakes, where fairly large seasonal fluctuations in
salinity can occur and where a single salinity measurement
may not be representative of the environment in which the
diatom lived. However, many calibration data sets are based
on limited water-chemistry sampling, often for logistical reasons. In such cases, problems can be minimized by sampling
a large number of sites and by making an effort to characterize
the range of chemical variability (Fritz et al., 1999).
Multivariate statistical techniques, such as canonical ordination (ter Braak, 1986; Birks, 1995), can then be used to
determine which environmental variables best explain the
distribution of diatom taxa. Environmental variables that
account for a high and significant proportion of the variation
in the assemblage can generally be used to develop quantitative inference models (called transfer functions), which are
then used to infer these variables from fossil assemblages.
The development of quantitative inference models
requires two steps: 1) a regression step, where species optima
and tolerances to a given variable are estimated, based on the
relative abundances of diatom taxa in the surface sediments
of lakes in the training set; and 2) a calibration step, where the
environmental variables are predicted from the relative abundances of diatom taxa in a fossil sample, based on the estimated species parameters obtained in the regression step. The
simplest and most statistically robust method proven so far
f o r deriving these transfer functions has been
weighted-averaging (WA) regression and calibration (Birks
et al., 1990), a method that is based on the observation that the
relationships between diatom taxa and environmental variables are often unimodal (Jongman et al., 1987; ter Braak and
van Dam, 1989). The logic is that, at a given value for a particular environmental variable (e.g. salinity), taxa with optima
nearest that value will be most abundant (ter Braak and
van Dam, 1989). A taxon's salinity optimum would then be
the average of all the salinity values for lakes in which the
taxon occurs, weighted by the taxon's abundance in each of
those lakes (the regression step). In the calibration step, the
estimated optima for the various species can then be used to
infer past salinity by taking an average of the species abundances in a fossil sample, each weighted by its WA optimum
(Birks et al., 1990).

S.E. Wilson and J.P. Smol

Because of the inherent circularity in generating


weighted-average inferences from the same data set used to
derive estimates of taxon optima and tolerances, prediction
error estimates based on the training set alone give overly
optimistic results (Birks et al., 1990). Computer-intensive
resampling procedures such as bootstrapping must therefore
be used to provide more realistic prediction errors associated
with the inference models (Birks et al., 1990; Birks, 1994;
Wilson et al., 1996).With bootstrapping, random samples are
drawn with replacement from the original data set to form a
bootstrap training set. Each sample in the original data set is
given an equal probability of selection, with the bootstrap
training set having as many samples as the original data set.
The samples that are not selected form the bootstrap test set.
In each bootstrap cycle (usually 1000 cycles in total),
weighted averaging regression and calibration are used to
estimate the optima and tolerances of taxa in the bootstrap
training set, and to infer the environmental variable in question for all sam~lesin the bootstrar, test set. This ~rocedureis
similar to using test sets in cross-validation, whereby a second independent test set of samples with actual measured
variables is used to infer the environmental variable in question; however, the bootstrap error estimate utilizes the full
size of the original data set rather than a smaller one, as in
cross-validation.
Fritz (1990) was the first to apply weighted-averaging
regression and calibration methods to derive a diatom-based
salinity transfer function. Developed from 64 largely saline
lakes of the northern Great Plains of Canada and the United
States, the transfer function was applied to a recent diatom
stratigraphy from Devils Lake, North Dakota, for which measured lake-water salinity values were available for the past
century. The diatom-inferred salinity values agree closely
with the measured values at the lower end of the salinity

spectrum (<lo g.L-l); however, there was more discrepancy


during high-salinity periods. These differences may relate to
the threshold response of diatom species to salinity change or
possibly to problems associated with the sediment record,
such as sediment reworking or poor diatom preservation
(Fritz, 1990; Gasse et al., 1997). The transfer function,
described by Fritz et al. (1993), has since been applied to
other longer cores from Devils Lake (Fritz et al., 1991; Fritz
et al., 1994) and to a core from Moon Lake, North Dakota
(Laird et a]., 1996a, b) to infer Holocene climatic change on
both decade and century time scales.
Recently, diatom-based salinity transfer functions have
been developed from saline lake regions in Africa (Gasse
et al., 1995), Spain (Reed, 1998a), and British Columbia
(Cumrning and Smol, 1993; Wilson et al., 1994, 1996), and
can be used to quantitatively infer historical changes in salinity from appropriately chosen closed-basin lakes with well
dated sediment cores.
In this paper, we briefly summarize the diatom data from
three Palliser Triangle lake sediment cores (Fig. 1): one from
Harris Lake, Cypress Hills, Saskatchewan that spans most of
the Holocene (Wilson et al., 1997), and two (an early
Holocene record and a recent sediment record) from
Clearwater Lake, Saskatchewan (Wilson, 1996; Last et al.,
1998; Leavitt et gl., 1999). preliminary analyses were done
on a number of other cores from Palliser Triangle lakes
(Chappice Lake, Alberta; Elkwater Lake, Cypress Hills,
Alberta; Kenosee Lake, Saskatchewan; Antelope Lake,
Saskatchewan), but poor diatom preservation precluded further study in almost every case. Although one nearshore core
from Elkwater Lake (EW3) appeared to have adequate diatom preservation in the upper 350 cm, the Harris Lake core
provided a much more extensive chronology and better

Figure 1. Location of Harris and Clearwater lakes within the Palliser Triangle (dashed line) and
Brown Chernozemic Soil Zone (shaded). Numbers denote other lakes that underwent preliminary
diatom analyses but had poor diatom preservation: 1 ) Chappice Lake, 2 ) Elkwater Lake, 3)Antelope
Lake, and 4) Kenosee Lake.

GSC Bulletin 534

Table 1 . Lake depth, surface area, and some


chemical measurements for Harris Lake and
Clearwater Lake. Salinity was calculated as the
sum of the major cations and anions listed below.
All chemistry values except pH are in m g L i and
are averages of spring, summer, and wintervalues.
Analyses provided by Bio-Chem Consulting
Services Ltd., Calgary, Alberta.
Variable

Harris Lake

Maximum depth (m)


Surface area
TDS
Salinity
Ca
Mg
Na
K

1.5
9 ha
230
338
39
52
14
2
19
235
0
4
2
8.2

so4
HCO,

co3
CI
Si
pH

Clearwater Lake

preservation, and was subsequently chosen as the more


appropriate Cypress Hills site. Antelope and Chappice lakes
had few to no diatoms in the surficial sediments (0 and 3 cm,
respectively) and core samples, while Kenosee Lake had well
preserved diatoms in the surficial sediments but substantially
degraded and much less abundant valves below 15 cm depth.
Surface sediment data sets from the northern Great Plains
(NGP)and British Columbia have shown that, in some cases,
poor preservation is related to very high salinity (>50 g . ~ - ' ) ,
often coupled with shallow lake depth, which may explain the
absence o f diatoms in Chappice Lake sediments. A recent
study o f diatom preservation in Spanish salt lakes, o f which
many were ephemeral, indicated that physical processes
accompanying lake desiccation played the most important
role in enhancing diatom dissolution, although high salinity
was also shown to be a factor for more permanent lakes
(Reed, 1998b).Explanations for the absence o f diatoms in the
recent sediments o f lower salinity lakes such as Kenosee and
Antelope are more difficult, although such cases are not
uncommon, particularly in carbonate-dominated, high-pH,
salt lakes (Cumming et al., 1995; S.E. Wilson, unpub. data,
1992,1994).Experimental dissolution of diatom valves in the
laboratory has shown that the dissolution rate is greatly
enhanced in Na2C03 solutions, where pH greater than 9 aids
the dissociation o f silicic acid (Barker et al., 1994).However,
neither Kenosee Lake nor Antelope Lake has a particularly
high carbonate concentration (Vance and Last, 1994).
Clearly, more research is required to determine other factors
involved in the dissolution process.
The following studies demonstrate some o f the successes
and difficulties encountered when using diatoms and other
paleolimnological indicators to reconstruct past climatic
changes from some topographically closed basin lakes, particularly those in regions with large anthropogenic influence,
such as the Palliser Triangle. It should be noted that, while

many o f the lakes in the Palliser Triangle are saline, Harris


Lake is currently fresh (<0.5 g . ~ - Table
l;
1 ) and Clearwater
;
1).
Lake is subsaline (0.5-3 g . ~ - l Table

METHODS
Field and laboratory methods, along with diatom enumeration protocols, are detailed in Wilson (1996)and Wilson et al.
(1997).Taxonomic identifications were based on a published
diatom flora from fresh and saline lakes in British Columbia
(Curnming et al., 1995),as well as Kramrner and Lange-Bertalot
(1986,1991a,b),Hikansson et al. (1993),and Carvalho et al.
(1995).
Paleosalinity reconstructions were generated from relative abundance data on the fossil diatom assemblages using a
salinity transfer function developed from 208 lakes in southern British Columbia and 11 lakes from the northern Great
Plains (Wilson et al., 1996). The 11 NGP lakes (six from
Saskatchewan) were added to the data set to compensate for a
lack o f Cyclotella choctawhatcheeana Prasad et al. in the
modern flora o f the British Columbia lakes, because this diatom was abundant in the early Holocene sediment record
from Clearwater Lake (see Wilson et al., 1996).This transfer
function was used in place o f the Fritz et al. (1993) NGP
transfer function for a number o f reasons, including 1 ) the
importance o f having taxonomic consistency between calibration and core samples; 2) the British Columbia data set
was larger and encompassed a broader range o f salinities,
particularly at the freshwater end, thus ensuring more accurate estimates o f diatom salinity optima and tolerances; and
3) the British Columbia transfer function was representative
o f most o f the taxa from the three sediment cores studied with
the exception o f C. clzoctawhatcheearza (Pl. 1,fig. 29,30) and
three other taxa from the early Holocene Clearwater Lake
core. Notably, these other three taxa appeared to be rare in the
modern flora o f the NGP lakes as well (Fritz et al., 1993;S.E.
Wilson, pers. observ., 1994). It should also be noted that,
although limnological differences exist between lakes o f
British Columbia and the NGP, including lake surface area,
depth, and, to some extent, ionic characterization, there are
some important similarities. For example, the saline lakes o f
both regions are largely dominated by Na-Mg-SO4,and many
have high concentrations o f phosphorus and nitrogen
(Campbell and Prepas, 1986; Cumming et al., 1995).
The computer program WACALIB v. 3.3 (Line et al.,
1994) was used to generate the salinity reconstructions for
each sediment record. T o assess the reliability o f individual
reconstructed salinity values in the early Holocene
Clearwater Lake core, a canonical correspondence analysis
(CCA),with axis 1 constrained to salinity, was performed on
the modern (those in the transfer function) and fossil diatom
samples, using the program CANOCO v . 3.1 (ter Braak,
1988, 1990). 'The squared residual distance o f each fossil
sample from the salinity axis was compared to the squared
residual distances o f the modern samples. Fossil samples
with residual distances equal to or greater than the extreme
5% o f the training set were considered to have a very poor fit
to the salinity axis (Birks et al., 1990). The estimated

S.E. Wilson and J.P. Smol

bootstrap standard error of prediction for each inferred salinity value was also plotted as a means of showing the reliability
of the reconstruction for the Harris Lake and Clearwater Lake
cores.

CASE STUDIES
Harris Luke (lat. 49 040.0iN, long. 109 '54.2' W, elev.
I234 m a.s.1.)
The diatom profile and salinity reconstruction of Harris Lake
are presented and discussed in detail in Wilson et al. (1997).
The salinity reconstruction shows that the lake remained
fresh (<0.5 g.L-') throughout the Holocene, with the highest
salinities occurring between approximately 6500 and
5200 BP (Fig. 2). During this time, salinity levels frequently
rose to about 0.2 or 0.3 g.L-l from values of about 0.1 g . ~ -or
l
less -fairly minor increases that are within the range of seasonal variation for Harris Lake (Wilson et al., 1997). This
period of slightly higher salinity corresponds to the only
period in the lake's history when the diatom flora was dominated by planktonic rather than benthic taxa. While the timing
of this change corresponds to the period of maximum
Holocene aridity, determined from ostracode (Porter, 1993)
and pollen studies (Sauchyn and Sauchyn, 1991) from Harris
Lake, it is unclear whether this shift in the diatom species
composition is a response to increased aridity. In
closed-basin lakes, increases in lake-water salinity generally
result from lowering of lake levels and subsequent concentration of cations and anions, coincident with a more arid climate. If the elevated salinities between 6500 and 5200 BP
represent the lowest water levels in the lake's history, it seems
unlikely that planktonic taxa would dominate only at this
time, especialiy if the lake was never more than 3 or 4 rn deep
at any point in its existence (Last and Sauchyn, 1993). As
such, this period is not considered to be one of especially low
water level. Another possibility is that this shift represents an
indirect response to climatic warming, whereby reduced surface cover in the watershed from a predominantly grassland
vegetation led to increased erosion and input of sediments
and nutrients to the lake (Sauchyn, 1990; Last and Sauchyn,
1993). The planktonic taxa present during this interval are
known to thrive in high-nutrient, eutrophic environments
(Wilson et al., 1997) and are able to persist at moderate abundances (>5%) in water depths as low as 3-4 m.
In general, the Holocene salinity levels inferred by the
diatoms agree with the ostracode data, which indicate that
salinities were never much higher than 0.45 g.L-l (Porter,
1993). However, the ostracode record indicates maximum
salinity levels prior to the middle Holocene and a reduction in
salinity between 6500 and 5200 BP (Porter, 1993),where diatom-inferred salinity was highest. Neither the diatom nor the
ostracode records indicate episodes of hypersalinity
(>50 g'L-l) between 9240 and 5500 BP, as documented
Last and Sauchyn (1993) based on mineralogical evidence.
However, because the diatoms and ostracodes were sampled
at 'Oarser intervals than the
(lo cm vs. approx'
5 cm), and different samples were used for each indicator, it is
possible that short-lived episodes were missed. Furthermore,

diatoms were not well preserved in the period before 6000 BP


(only about five samples had diatoms), making comparisons
particularly difficult for this period. Whether sampling resolution can explain all of the observed discrepancies between
the diatom and mineralogical records is difficult to say without further detailed analyses using the same samples. Clearly,
the existence of freshwater, planktonic and benthic diatoms
between 6500 and 5200 BP, along with periodic occurrences
of gypsum and thenardite, requires a shift from freshwater
conditions (<0.5 g . ~ - l )to the highly saline environment
required for precipitation of these salts (40 g-L-Iand 80 g . ~ - l ,
respectively) within a relatively short period of time (approx.
10 years or less). Although there is a remote possibility that
the salts are of authigenic origin, the most likely scenario for
their formation requires at least partial drying of the lake
basin (W. Last, pers. cornm., 1995), whereas the diatom data
do not support periodic drying during this time.
Overall, the diatoms, ostracodes, and mineralogy indicate
that Harris Lake was a freshwater basin throughout most of its
history, including much of the early Holocene. The relatively
unchanged diatom-inferred salinity record indicates a low
sensitivity to changes in precipitation and evaporation compared to other closed-basin lakes in the Great Plains region.
This is almost certainly the result of groundwater dominance
and a relatively minor contribution from direct precipitation
and surface runoff (Last and Sauchyn, 1993). Unfortunately,
the lack of a complete diatom record, particularly during the
early Holocene, precludes a full comparison with climate
records from other sites on the lower plains, particularly with
respect to the timing of early Holocene warming in upland
versus lowland sites. As such, this study demonstrates the
importance of using multiple indicators when interpreting
past climatic and environmental changes.

Clearwater Luke (lat. 5052.3'N, long. 107O55.8'W,


elev, 680 a,s.~.)
Early Holocene record
The early Holocene diatom record from Clearwater Lake is
discussed in detail in Wilson (1996) and Last et al. (1998).
The diatom-inferred salinity indicates that the basin experienced broad fluctuations in salinity (approx. 2-20 g . ~ - l )
between 9700 and 9400 BP, and was fully saline by 9400 BP
(Fig. 3). Maximum salinity concentrations at Cleanvater
Lake were attained during the period 9650-9400 BP and in
the interval 9200-8600 BP. However, the lake may have
reached salinities of this magnitude later in the Holocene as
well. By 8200 BP, the lake had begun a transformation to a
subsaline or freshwater lake. At some point subsequent to
6600 BP, the water level in the basin dropped and the remaining Holocene record at the core site was lost through either
nondeposition or erosion (Last et al., 1998).
Modern analogues for the diatom assemblages present
between 9400 and 7900 BP were often scarce (Fig. 3). This
was largely due to the abundance of Achnanthes thermalis
(PI. 1, fig. 18,19), Navicula aff. salinicola (PI. 1, fig. 17), and
Amphora aff coffeaformis, which were not well represented
in the modern calibration lakes used to develop the salinity

GSC Bulletin 584

PLATE 1
Photomicrographsand one scanning electron micrograph (fig. 30) of some of the common
diatoms found in the Harris and Clearwater lake cores. Scale bar on figure 1 is 10 pm and
applies to all figures except 5 (where scale bar = 10 pm) and 30 (wherescale bar = 2pm).
Figure 1.
Figure 2.
Figure 3.
Figure 4.
Figure 5.
Figure 6.
Figure 7.
Figure 8.
Figure 9.
Figure 10.
Figure 11.
Figure 12.
Figure 13.
Figure 14.
Figure 15.
Figure 16.
Figure 17.
Figure 18.
Figure 19.
Figure 20.
Figure 21.
Figure 22.
Figure 23.
Figure 24.
Figure 25.
Figure 26.
Figure 27.
Figure 28.
Figure 29.
Figure 30.

Stephanodiscus niagarae Ehrenb.


Mastogloia smithii Thwaites.
Mastogloia smithii Thwaites.
Epithemia argus (Ehrenb.) Kutz.
Navicula oblonga (Kutz) Kutz. Note change in scale (scale bar = 10 pm).
Cymbella pusilla Grun. ex A. Schmidt.
Denticula valida (Ped.) Grun. in V.H.
Denticula valida (Ped.) Grun. in V.H.
Denticula kuetzingii Grun.
Amphora libyca Ehrenb.
Cocconeis placentula var. euglypta (Ehrenb.) Grun.
Brachysira aponina Kutz.
Navicula bulnheimii Grun. in V. H.
Cymbella microcephala Grun. in V.H.
Cymbella sp. 1 PISCES.
Amphora pediculus (Kutz) Grun. ex A. Schmidt.
Navicula aff , salinicola.
Achnanthes thermalis (Raben.) Schoenfeld.
Achnanthes thermalis (Raben.) Schoenfeld.
Achnanthes minutissima Kutz.
Fragilaria brevistriata Grun. in V.H.
Fragilaria construens var. venter (Ehrenb.) Grun, in V.H.
Fragilaria pinnata Ehrenb.
Fragilaria construens var. venter (Ehrenb.) Grun. in V.H.
Fragilaria construens (Ehrenb.) Grun.
Chaetoceros muelleri Lemm. cysts.
Chaetoceros muelleri Lemm. cysts.
Aulacoseira ambigua (Grun. in V.H.) Simonsen.
Cyclotella choctawhatcheeanaPrasad et al.
Cyclotella choctawhatcheeana Prasad et al. Note change in scale (scale bar = 2 pm).

S.E. Wilson and J.P. Smol

Plate 1

GSC Bulletin 534

transfer function (Wilson et al., 1996). A. thermalis was not


found in the surface sediments of any of the 224 lakes sampled from British Columbia (Wilson et al., 1996) or the 64
lakes sampled from the NGP (Fritz et al., 1993),while both N.
aff. salinicola and A. aff. coffeaeformis were present, but only
occasionally and at very low abundance. Salinity reconstructions should therefore be considered tentative for some intervals, although it is quite likely, in these instances, that
salinities were higher than indicated in Figure 3 (Last et al.,
1998).

Moose Jaw, Saskatchewan (Yansa, 1998) indicate that the


transition to fully saline conditions occurred much later in the
Holocene (ca. 8000-6000 BP). Likewise, at many of these
sites, this high-salinity phase persisted considerably longer
into the Holocene than is evident at Clearwater Lake,
although the truncated record does not permit a complete
comparison. Evidence from pollen and diatoms for late glacial-early Holocene high-salinity episodes is also found in
central Alberta (Hickman and Schweger, 1993), although the
chronologies of these records are somewhat tentative.

The early Holocene high-salinity phase corresponds well


with a number of other late glacial-early Holocene lacustrine
records from the region. Radle et al. (1989) and Kennedy
(1994) indicated that Medicine Lake in South Dakota
changed from a freshwater lake at around 10 000 BP to a
highly saline basin at 9500 BP. Similarly, the geochemistry of
ostracode shells and endogenic inorganic carbonates from
Coldwater Lake in North Dakota shows that an increase in
salinity occurred about 9000 BP (Xia et al., 1997).

At most of the above sites, high salinities during the early


to middle Holocene have been interpreted as periods of lower
water level resulting from evaporative concentration under a
warmer andlor drier climate. This warm period has been
explained by increased summer insolation at high latitudes
between 12 000 and 9000 BP, induced by Milankovitch
eartwsun orbital variations (Barnosky et al., 1987; Schweger
and Hickman, 1989; Ritchie and Harrison, 1993), and generally occurred somewhat later in the midwestern and eastern
parts of the continent where the climate was controlled for a
longer period by the retreating Laurentide Ice Sheet

In contrast, sediment records from Devils Lake in North


Dakota (Fritz et al., 1991;Haskell et al., 1996),Moon Lake in
North Dakota (Laird et al., 1996a), Ceylon Lake in
Saskatchewan (Last, 1990), and the Andrews dugout near

7000 1

I I Fit

1470+40BPnI*

100 200 3450 + 50 BPI*

300 400 -

-56

500 5120+60BPn*

600 700 800 -

Mazarna ash
ca. 6800 BP

'

900

Diatom-inferred salinity ( g - ~ - ' )


Diatom-inferred salinity (g.~-l)
Figure 2. Plot of the diatom-inferred salinity for Harris Lake.
The inferred salinity values are mean bootstrap estimates and
the errors (dotted lines) are estimated standard errors of
prediction. Chronological control is provided by four
radiocarbon dates plotted on the y-axis, as well as by the
occurrence of Mazama tephra (ca. 6800 BP).

Figure 3. Plot of the diatom-inferred salinity for Cleanvater


Lake (ca. 9700-7100 BP). Five diatom zones are labelled
CWDl to CWD.5. Solid circles on the Fit axis indicate fossil
diatom assemblages that were not well represented in the
modern diatom assemblages of the calibration lakes (see text
for details). The inferred salinity values are mean bootstrap
estimates and the errors (dotted lines) are estimatedstandard
errors of prediction. Six AMS radiocarbon dates provide the
chronology for the core (see Wilson (1996)for details).

S.E. Wilson and J.P. Smol

(Barnosky et al., 1987). The fact that the highest salinities at


Cleanvater Lake occurred prior to ca. 8700 BP probably does
reflect this early Holocene maximum summer insolation.
However, the fairly abrupt termination of the high-salinity
phase and formation of a freshwater lake after 8400 BP may
indicate an increased influence of groundwater flow at this
time, although there is a possibility that this event was climatically induced (Last et al., 1998).
Approximately AD 1600 to present
The diatom stratigraphy and diatom-inferred salinity of
Clearwater Lake from ca. AD 1600 to the present are discussed in detail in Wilson (1996) and Leavitt et al. (1999).
The diatom-inferred salinity indicates that Clearwater Lake
fluctuated within a narrow range of salinity during the past
400 years, remaining within the subsaline (0.5-3 g . ~ - l to
)

freshwater (<0.5 g . ~ - l range


)
(Fig. 4), similar to conditions
that prevailed around 7100 BP (Fig. 3). Slight increases in
salinity occurred ca. 1580-1620, ca. 1710-1750, ca.
1920-1940, and, most recently, ca. 1993 (Fig. 4).
Striking changes in the composition of the diatom community occurred ca. 1922, and also ca. 1977. While some of
the changes ca. 1922, such as the appearance of the
hyposaline Cymbella pusilla Grun, ex A. Schmidt (salinity
optimum = 5.47 g.L-l; Pl. 1, fig. 6; Wilson et al. (1996)), suggested a shift to more shallow, saline conditions, the diatom-inferred salinity remained relatively unchanged,
indicating that these shifts in community composition were
largely unrelated to changes in lake-water salinity. Similarly,
shifts in the diatom community that occurred ca. 1977 and in
the mid-1980s did not produce notable changes in the diatom-inferred salinity, despite multiyear drought in 19871989 (Jones, 1991) and known lake-level decline.
Overall, the diatom data suggest that Clearwater Lake has
not been especially sensitive to droughts during the last century. For example, severe droughts on the northern Great
Plains during 1860 (Bark, 1978; Stockton and Meko, 1983)
and in the early 1890s (Borchert, 1971; Bark, 1978; Jones,
1987) were not marked by an increase in diatom-inferred
salinity. Well known droughts of the 1930s and 1980s also
resulted in few changes in diatom-inferred salinity, despite
observed lake-level declines at Clearwater Lake during these
periods. As such, Cleanvater Lake is not very sensitive to
short-term climatic fluctuations when compared with other
lakes of the northern Great Plains (e.g. Fritz, 1990; Laird
et al., 1996a, b).
The fact that the most dramatic shifts in the diatom community do not correspond to well defined changes in salinity
suggests that other factors have had an overriding influence.
The shift ca. 1922 corresponds well with the major influx of
European settlers (1906-1921; Jones, 1987). Similarly, the
shift in the assemblage after ca. 1977 may relate to cottage
development around the lake. Cottages were almost nonexistent in 1940, but encircled all but the north shore of the lake by
1979. Catchment disturbances, including changes in land use,
are known to increase erosional and nutrient inputs, and may
have had a fertilizing effect on diatom assemblages
(Anderson et al., 1995).

0.05

0.1

0.2

0.5

Diatom-inferred salinity (gC1)

SUMMARY AND GENERAL DISCUSSION

The Harris Lake diatom record challenges some of the


previous interpretations of the early to middle Holocene cliFigure 4. Plot of the diatom-inferred salinity for Cleanvater
hate history of the Cypress Hills; however, apparent discrepLake (ca. AD 1600 to present). The inferred salinity v a l ~ ~ e s ancies with the mineralogical record, particularly in the early
are mean bootstrap estimates and the errors (dotted lines)
Holocene, could be the result of poor diatom preservation or
are estimated standard errors ofprediction. The chronology
differences in sampling resolution. In general, the diatoms,
of the core is based on 2 1 0 ~dating
b
of a duplicate core of tlze
mineralogy, sedimentology, and ostracodes indicate that
same length, using the constant rate of sz~pply(CSR) model
Harris Lalte was a freshwater basin throughout much of its
(Appleby and Oldfield, 1978). Further details are available
history, including the early Holocene, with the mineralogy
indicating short-lived episodes of hypersalinity in the early to
in Turner (1994),Last et al. (1998),and Leavitt et al. (1999).

GSC Bulletin 534


middle Holocene. This differs from most other early to middle Holocene sediment records from the northern Great
Plains, and from a number of currently freshwater lakes in
east-central Alberta (Schweger and Hickman, 1989;
Hickrnan and Schweger, 1993) that indicate higher salinities
for some time during this period (see also Vance et al., 1992).
The Clearwater Lake record also indicates an approximately
1000-year period of high salinity between ca. 9600 and
8600 BP, despite having a modern-day salinity close to that of
freshwater. The difference in early Holocene salinity
between Harris and Clearwater lakes probably indicates a
substantial amount of groundwater input to Harris Lake, even
at a time when the climate was much warmer and drier. The
records might also suggest that the warmer and drier climate
in the early to middle Holocene experienced by many lakes in
the NGP may not have been as pronounced in some elevated
areas such as the Cypress Hills. To some extent, this would
also relate to local hydrological factors. However, the
absence of other complete Holocene records from the
Cypress Hills and other higher elevation areas such as the
Moose Mountain upland, combined with the paucity of lakes
in these areas, makes it difficult to resolve this issue.
The low sampling resolution in the upper metre of the
Harris Lake core precludes compai-isons with the recent sediment record from Clearwater Lake, and with other
high-resolution studies of the last approximately 2000 years
from lakes in the NGP (e.g. Vance et al., 1992, 1993; Fritz
et al., 1994; Laird et al., 19966, 1998). However, it is clear
from the diatom, ostracode, and mineralogical records of
Harris Lake that hydrological conditions remained quite stable throughout the latter half of the Holocene. For example,
the sediment record gives no indication of fluctuating water
level and/or salinity associated with the Medieval Warm
Period and Little Ice Age, as recorded in other NGP lakes
such as Chappice Lake in Alberta (Vance et al., 1993), Moon
Lake in North Dakota (Laird et al., 1996a, b), Kenosee Lake
in Saskatchewan (Vance et al., 1997), Redberry Lake in
Saskatchewan (Van Stempvoort et al., 1993), and Waldsea
Lake in Saskatchewan (Last and Slezak, 1988). Once again,
the dominance of groundwater in the hydrological budget of
Harris Lake is evident from the almost uniform salinities
during the past 4000 years. This should not be too surprising,
considering that Harris Lake is presently fresh. However, preliminary analyses at Elkwater Lake, a freshwater lake in the
Cypress Hills to the west of Harris Lake, indicate a change
from a lower water level and hyposaline conditions to a
higher water level and freshwater conditions around the
mid-Holocene (Vance and Last, 1994). At present it is not
clear whether this change is the result of local hydrological or
climatic factors.
Diatom analyses from the recent sediment record of
Clearwater Lake indicate that factors other than salinity were
responsible for dramatic changes in the diatom assemblage
over the past 400 years. Analysis of multiple proxy data indicates that land-use changes in the Clearwater Lake catchment
were almost certainly responsible for these changes, despite

documented water-level changes and known droughts. These


findings suggest that basin-specific factors, such as human
land-use history and hydrological setting, can moderate the
effects of climate on lake systems (e.g. Van Stempvoortet al.,
1993).The fact that Clearwater Lake is presently almost fresh
(approx. 1 g . ~ - l )together
,
with the significant annual moisture deficit of the NGP, suggests that the water chemistry of
Clearwater Lake may be buffered against drought due to fresh
groundwater inputs similar to those at Harris Lake. The early
Holocene record from Clearwater Lake indicates that relatively freshwater conditions (<I g . ~ - lprevailed
)
by 7900 BP,
although the truncated record makes it impossible to say
whether the lake remained fresh through the mid- to late
Holocene. However, the fact that the lake is currently fresh
lends some support to the hypothesis that the groundwater-dominated system established very early on in the
Clearwater Lake basin remained through to the present day.
The above studies demonstrate that diatom-based salinity
transfer functions are useful for inferring salinity changes
from lake sediment cores. However, problems do exist, such
as poor diatom preservation at some sites and the absence of
good modern analogues for some fossil diatom assemblages.
Ideally, as more modern lake environments in the NGP are
studied, the number of potential analogues for sediment core
diatom assemblages should increase. However, we have to
consider that these problems may not be easy, or even possible, to resolve with expansion of data sets, simply because
modern analogues may no longer exist. For example, high
nutrient loading to many saline lakes, because of their location in areas of intense agriculture and other anthropogenic
influences, makes it difficult to find analogues for highly
saline, low-nutrient environments that may have existed in
the past. Alternatively, analogues may exist in hydrological
regions outside western ~ a n a d a(e.g. Africa, ~ustralia),
pointing to the importance of merging regional calibration
data sets into larger, global ones (see Juggins et a]., 1994;
Fritz et al., 1999).
Finally, the above case studies have shown that catchment
disturbances and local hydrological setting, in particular, are
important factors regulating responses in closed-basin lakes,
and may obscure climate signals. Ideally, lakes chosen for
paleoclimatic studies should be those that are known to have
responded to droughts in the recent past and where diatom-inferred salinity has been successful at tracking changes
in the instrumental climate record (e.g. Laird et al., 1996a).
Admittedly, locating such sites in the Palliser Triangle has
proved difficult (Vance and Last, 1994), particularly those
with continuous Holocene sediment records. However. one
site that looks very promising is Oro Lake (Vance and ~ a s t ,
1996), a climatically sensitive, saline lake located on The
Missouri Coteau of southern Saskatchewan. By virtue of the
9500-year long, finely laminated sediment record that has
been retrieved from the centre of the basin, Oro Lake may be
one of the more important lacustrine sequences yet retrieved
from the Great Plains region of western Canada (Vance and
Last, 1996).

S.E. Wilson and J.P. Smol

ACKNOWLEDGMENTS
This project was funded by a Natural Sciences and Engineering Research Council of Canada (NSERC) strategic grant to
JPS, and by the Geological Survey of Canada Palliser
Triangle IRMA. We thank W.M. Last, D.S. Lemmen,
D.J. Sauchyn, and R.E. Vance for help with various aspects of
this research, and D.S. Lemmen, R.E. Vance, and C. PrCvost
for their very helpful comments on the manuscript.

REFERENCES
Anderson, N.J., Renberg, I., and Segerstrom, U.
1995: Diatom production responses to the development of early agriculture in a boreal forest lake-catchment (Kassjon, northern Sweden);
Ecology, v. 83, p. 809-822.
Appleby, P.G. and Oldfield, F.
1978: Thecalculation of lead-210 dates assuming a constant rate of supply
of unsupported 210Pb to the sediment; Catena, v. 5, p. 1-8.
Bark, L.D.
1978: History of American droughts; in North American Droughts, (ed.)
N.J. Rosenberg; American Association for the Advancement of
Science, Selected Symposium 15, Westview Press Inc., Boulder,
Colorado, p. 9-23.
Barker, P., Fontes, J., Gasse, F., and Druart, J.
1994: Experimental dissolution of diatom silica in concentrated salt solutions and implications for paleoenvironmental reconstruction;
Limnology and Oceanography, v. 39, p. 99-1 10.
Barnosky, C.W., Grimm, E.C., and Wright, H.E., Jr.
1987: Towards apostglacial history of the northern Great Plains: areview
of the paleoecologic problems; Annals of the Carnegie Museum,
V. 56, p. 259-273.
Birks, H.J.B.
1994: Theimportanceof pollen and diatom taxonomic precision in quantitative palaeoenvironmental reconstructions; Review of
Palaeobotany and Palynology, v. 83, p. 107-117.
1995: Quantitative palaeoenviron~nentalreconstructions; in Statistical
Modelling of Quaternary Science Data, (ed.) D. Maddy and
J.S. Brew; Quaternary Research Association, Cambridge, United
Kingdom, Technical Guide 5,271 p.
Birks, H.J.B., Line, J.M., Juggins, S., Stevenson A.C.,
and ter Braak C.J.F.
1990: Diatoms and pH reconstruction;Philosophical Transactions, Royal
Society of London, Series B, v. 327, p. 263-278.
Blim, D.W.
1993: Diatom community structure along physicoche~nicalgradients in
saline lakes; Ecology, v. 74, p. 1246-1263.
Borchert, J.R.
1971: The dust bowl in the 1970s; Annals of the Association of American
Geographers,v. 61, p. 1-22.
Bradbury, J.P., Forester, R.M., and Thompson, R.S.
1989: Late Quaternary paleolimnology of Walker Lake, Nevada; Journal
of Paleolimnology, v. 1, p. 249-267.
Campbell, C.E. and Prepas, E.E.
1986: Evaluation of factors related to the unusually low chlorophyll levels
in prairie saline lakes; Canadian Journal of Fisheries and Aquatic
Sciences, v. 43, p. 846-854.
Carvalho, L.R., Cox, E.J., Fritz, S.C., Juggins, S., Sims, P.A.,
Gasse, F., and Battarbee, R.W.
1995: Harmonising the taxonomy of salt lake diatoms: Cyclotella caspia
as a case study; Diatom Research, v. 10, p. 229-240.
Charles, D.F. and Smol, J.P.
1994: Long-term chemical changes in lakes: quantitative inferences using
biotic remains in the sediment record; in Environmental Chemistry
of Lakes and Reservoirs, (ed.) L. Baker; American Chemical
Society, Washington, D.C., Advances in Chemistry Series 237,
p. 3-31.
Cumming, B.F. and Smol, J.P.
1993: Development of diatom-based models for paleoclimatic research
from lakes in British Columbia (Canada); Hydrobiologia,
v. 2691270, p. 179-196.

Cumming, B.F., Wilson, S.E., Hall, R.I., and Smol, J.P.


1995: Diatoms from British Columbia (Canada) and their relationship to
salinity, nutrients, and other limnological variables; Bibliotheca
Diatomologica Band 31, Gebriider Borntraeger, BerlinIStuttgart,
Germany, 207 p.
Dixit, S.S., Cumming, B.P., Birks, H.J.B., Smol, J.P., Kingston, J.C.,
Uutala, A.J., Charles, D.F., and Camburn, K.E.
1993: Diatom assemblagesfrom Adirondack lakes (New York, USA) and
the developmentof inference models for retrospective environmental assessment; Journal of Paleolimnology, v. 8, p. 2 7 4 7 .
Dixit, S.S., Smol, J.P., Kingston, J.C., and Charles, D.F.
1992: Diatoms: powerful indicators of environmental change; Environmental Science and Technology, v. 26, p. 22-32.
Fritz, S.C.
1990: Twentieth-century salinity and water-level fluctuations in Devils
Lake, North Dakota: test of a diatom-based hansfer function;
Limnology and Oceanography, v. 35, p. 1771-1781.
Fritz, S.C. and Battarbee, R.W.
1988: Sedimentary diatom assemblages in freshwater and saline lakes of
the northern Great Plains, North Ame~.ica:preliminary results; in
Proceedings of the 9th International Diatom Symposium, (ed.)
F.E. Round; Koeltz Scientific Books, Koenigstein, Germany,
p. 265-27 1.
Fritz. S.C.. Cumming. B.F.. Gasse F.. and Laird, K.R.
s i2icato;s of hydrologic and climatic change in saline
1999:. ~ k t o m as
lakes; in The Diatoms, (ed.) E. Stoermer and J.P. Smol;
Applications for the Environmental and Earth Sciences, Cambridge
University Press, Cambridge, United Kingdom, p. 41-72.
Fritz, S.C., Engstrom, D.R., and Haskell, B.J.
1994: 'Little Ice Age' aridity in the North American Great Plains: a high
resolution reconstruction of salinity fluctuations from Devils Lake,
North Dakota, USA; The Holocene, v. 4, p. 69-73.
Fritz, S.C., Juggins, S., and Battarbee, R.W.
1993: Diatom assemblages and ionic characteri7ation of freshwater and
saline lakes of the northern Great Plains, North America: a tool for
reconstructing past salinity and climate fluctuations; Canadian
Journal of Fisheries and Aquatic Sciences, v. 50, p. 1844-1856.
Fritz, S.C., Juggins, S., Battarbee, R.W., and Engstrom, D.R.
1991: Reconstruction of past changes in salinity and climate using a
diatom-based transfer function; Nature, v. 352, p. 706-708.
Gasse, F.
1987: Diatoms for reconstructing palaeoenvironments and
paleohydrology in tropical semi-arid zones: example of some lakes
from Niger since 12 000 BP; Hydrobiologia, v. 154, p. 127-163.
Gasse, F., Fontes, J.C., Plaziat, J.C., Carbonel, P., Kaczmarska, I.,
De Deckker, P., SouliB-Marsche, I., Callot Y., and Dupeuble, P.A.
1987: Biological remains, geochemistry and stable isotopes for the reconstruction of environmental and hydrological changes in the
Holocene lakes from North Sahara; Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 60, p. 1 4 6 .
Gasse, F., Barker, P., Gell, P., Fritz, S.C., and Chali&,F.
1997: Diatom-inferred salinity in palaeolakes: an indirect tracer of climate
change; Quaternary Science Reviews, v. 16, p. 547-563.
Gasse, F., Juggins, S., and Ben Khelifa, L.
1995: Diatom-based transfer functions for inferring past hydrochemical
characteristics of African lakes; Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 117, p. 31-54.
Gasse, F., Talling, J.F., and Kilham, P.
1983: Diatom assemblages in East Africa: classification, distribution,and
ecology; Revue dlHydrobiologie Tropicale, v. 16, p. 3-34.
Gell, P.A. and Gasse, F.
1994: Relationships between salinity and diatom flora from some Australian saline lakes; in Proceedings of the l l th International Diatom
Symposium, San Francisco, California, (ed.) J.P. Kociolek;
California Academy of Sciences, San Francisco, p. 631-647.
Gell, P.A., Barker, P.A., De Deckker, P., Last W.M., and Jelicic, L.
1994: The Holocene history of West Basin Lake, Victoria, Australia:
chemical changes based on fossil biota and sediment mineralogy;
Journal of Paleolimnology, v. 12, p. 235-258.
Hhkansson, H., Hajdu, S., Snoeijs, P., and Loginova, L.
1993: Cyclotella hakanssoniae Wendker (Bacillariophyceae), and its
relationship to C. caspia Grunow and other similar brackish water
Cyclotella species; Diatom Research, v. 8, p. 333-347.

GSC Bulletin 534


Hall, R.I. and Smol, J.P.
1992: A weighted-averaging regression and calibration model for inferring total phosphorus concentration from diatoms in British
Columbia (Canada) lakes; Freshwater Biology, v. 27, p. 417-434.
Hammer, U.T.
1990: The effects of climate change on the salinity, water levels and biota
of Canadian ~rairiesaline lakes: Verhandlungen der Internationalen
~ e r e i n i ~ uvon
n g Limnologen,". 24, p. 3211326.
Hasltell, B.J., Engstrom, D.R., and Fritz, S.C.
1996: Late Quaternary paleohydrology in the North American Great
Plains inferred from the geochemistry of endogenic carbonate and
fossil ostracodes from Devils Lake, North Dakota, USA; Palaeogeography, Palaeoclimatology,Palaeoecology, v. 123, p. 161-178.
Hickman, M. and Schweger, C.E.
1993: Late glacial-early Holocene palaeosalinity in Alberta, Canada climate implications;Journal of Paleolimnology,v. 8,p. 149-161.
Jones, D.C.
1987: Empire of Dust; University of Alberta Press, Edmonton, Alberta,
316 p.
Jones, K.H.
1991: Drought on the Prairies; in Symposium on the Impacts of Climate
G. Wall; UniverChange and Variability on the Great Plains, (4.)
sity of Waterloo, Waterloo, Ontario, Department of Geography,
Occasional Paper No. 12, p. 125-129.
Jongman, R.H.G., ter Braak, C.J.F., and van Tongeren, O.P.R.
1987: Data Analysis in Community and Landscape Ecology; Pudoc,
Wageningen, The Netherlands, 299 p.
Juggins, S., Battarbee, R.W., Fritz S.C., and Gasse, F.
1994: The CASPIA project: diatoms, salt lakes, and environmental
change; Journal of Paleolimnology, v. 12, p. 191-196.
Kennedy, K.A.
1994: Early Holocene geochemical evolution of saline Medicine Lake,
South Dakota; Journal of Paleolimnology, v. 10, p. 69-84.
Kingston, J.C. and Birks, H.J.B.
1990: Dissolved organiccarbon reconstructions from diatom assemblages
in PIRLA project lakes, North America; Philosophical Transactions, Royal Society of London, Series B, v. 327, p. 279-288.
Krammer, K. and Lange-Bertalot, H.
1986: Bacillariophyceae. I. Teil: Naviculaceae; in SiiRwasserflora von
Mitteleuropa, Band 211, (ed.) H. Ettl, G. Gartner, J. Gerloff,
H. Heynig, and D. Mollenhauer; Gustav Fischer Verlag, Stuttgart,
Germany, 876 p.
1991a: Bacillariophyceae.3. Teil: Centrales, Fragilariaceae, Eunotiaceae;
in SiiBwasserflora von Mitteleuropa, Band 213, (ed.) H. Ettl,
G. Gartner, J. Gerloff, H. Heynig, and D. Mollenhauer; Gustav
Fischer Verlag. Stuttgart, Germany, 576 p.
1991b: Bacillariophyceae. 4. Teil: Achnanthaceae Kritische Erganzungen
zu Navicula (Lineolateae) und Gomphonema; in SiiBwasserflora
von Mitteleuropa, Band 214, (ed.) H. Ettl, G. Gartner, J. Gerloff,
H. Heynig, and D. Mollenhauer; Gustav Fischer Verlag, Stuttgart,
Germany, 437 p.
Laird, K.R., Fritz, S.C., and Cumming, B.F.
1998: A diatom-based reconstruction of drought intensity, duration, and
frequency from Moon Lake, North Dakota: a sub-decadal record of
the last 2300 years; Journal of Paleolimnology,v. 19, p. 161-179.
Laird, K.R., Fritz, S.C., Grimm, E.C., and Mueller, P.G.
1996a: Century-scale paleoclimatic reconsbvction from Moon Lake, a
closed-basin lake in the northern Great Plains; Limnology and
Oceanography, v. 41, p. 890-902.
Laird, K.R., Fritz, S.C., Maasch, K.A., and Cumming, B.F.
1996b: Greater drought intensity and frequency before AD 1200 in the
northern Great Plains, USA; Nature, v. 384, p. 552-554.
Last, W.M.
1990: Paleohydrology and paleochemistry of Ceylon Lake, Canada;
Journal of Paleolimnology, v. 4, p. 219-228.
Last, W.M. and Sauchyn, D.J.
1993: Mineralogy and lithostratigraphy of Harris Lake, southwestern
Saskatchewan,Canada; Journal of Paleolimnology,v. 9, p. 23-39.
Last, W.M. and Slezak, L.A.
1988: The salt lakes of western Canada: a paleolimnological overview;
Hydrobiologia, v. 158, p. 301-316.
Last, W.M., Vance, R.E., Wilson, S.E., and Smol, J.P.
1998: A multi-proxy record of rapid early Holocene hydrologic change on
the northern Great Plains of southwestern Saskatchewan, Canada;
The Holocene, v. 8, p. 503-520.

Leavitt, P.R., Vinebrooke, R.D., Hall, R.I., Wilson, S.E., Smol, J.P.,
Vance, R.E., and Last, W.M.
1999: Multiproxy record of prairie lake response to climate change and
human activity, Clearwater Lake, Saskatchewan; in Holocene
Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Line, J.M., ter Braak, C.J.F., and Birks, H.J.B.
1994: WACALLB version 3.3 - a computer program to reconstruct environmental variables from fossil assemblagesby weighted averaging
and to derive sample-specific errors of prediction; Journal of
Paleolimnology, v. 10, p. 147-152.
Metcalfe, S.E.
1988: Modern diatom assemblages in central Mexico: the role of water
chemistry and other environmental factors as indicated by
TWINSPAN and DECORANA; Freshwater Biology, v. 19,
p. 217-233.
Patrick, R. and Reimer, C.
1966: The Diatoms of the United States Exclusive of Alaska and Hawaii.
Volume 1. Fragilariaceae, Eunotiaceae, Achnanthaceae,
Naviculaceae; The Academy of Natural Sciences of Philadelphia,
Monograph 13,688 p.
Pienitz, R. and Smol, J.P.
1993: Diatom assemblages and their relationship to environmental variables in lakes near Yellowknife (N.W.T.,Canada); Hydrobiologia,
v. 2691270, p. 391-404.
Pienitz, R., Smol J.P., and Birks, H.J.B.
1995: Assessment of freshwater diatoms as quantitative indicators of past
climatic change in the Yukon and Northwest Territories, Canada;
Journal of Paleolimnology, v. 13, p. 2149.
Pienitz, R., Walker, I.R., Zeeb, B.A., Smol J.P., and Leavitt, P.R.
1992: Biomonitoring past salinity changes in an athalassic subarctic lake;
International Journal of Salt Lake Research, v. 1, p. 91-123.
Porter, S.C.
1993: A reconstruclion of Holocene environments based on ostracodes,
Harris Lake watershed, Cypress Hills, Saskatchewan; M.Sc, thesis,
University of Regina, Regina, Saskatchewan, 112 p.
Radle, N., Keister, C.M., and Battarbee, R.W.
1989: Diatom, pollen, and geochemical evidence for the palaeosalinity of
Medicine Lake, South Dakota, during the Late Wisconsin and early
Holocene; Journal of Paleolimnology, v. 2, p. 159-172.
Reavie, E.D., Hall, R.I., and Smol, J.P.
1995a: An expanded weighted-averaging model for inferring past total
phosphorus concentrations from diatom assemblages in eutrophic
British Columbia (Canada) lakes; Journal of Paleolimnology,v. 14,
p. 49-62.
Reavie, E.D, Smol, J.P., and Carmichael, N.B.
1995b: Postsettlement eutrophication histories of six British Columbia
(Canada) lakes; Canadian Journal of Fisheries and Aquatic
Sciences, v. 52, p. 2388-2401.
Reed, J.M.
1998a: A diatom-conductivity transfer function for Spanish salt lakes;
Journal of Paleolimnology, v. 18, p. 399-416
1998b: Diatom preservation in the recent sediment record of Spanish saline
lakes: implications for palaeoclimate study; Journal of
Paleolimnology, v. 18, p. 129-137.
Remenda, V.H. and Birks, S.J.
1999: Groundwater in the Palliser Triangle: an overview of its vulnerability and potential to archive climate information; in Holocene
Cl~mateand Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Ritchie, J.C. and Harrison, S.P.
1993: Vegetation, lake levels, and climate in western Canada during the
Holocene; in Global Climates Since the Last Glacial Maximum,
(ed.) H.E. Wright, Jr., J.E. Kutzbach, T. Webb, HI, W.F. Ruddiman,
F.A. Street-Perrott, and P.J. Bartlein; University of Minnesota
Press, Minneapolis, Minnesota, p. 401414.
Sauchyn, D.J.
1990: A reconstructionof Holocene geomorphology and climate, western
Cypress Hills, Alberta and Saskatchewan; Canadian Journal of
Earth Sciences, v. 27, p. 1504-1510.

S.E. Wilson and J.P. Smol


Sauchyn, M.A. and Sanchyn, D.J.
1991: A continuous record of Holocene pollen from Harris Lake, southwestern Saskatchewan, Canada; Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 88, p. 13-23.
Schweger, C.E. and Hickman, M.
1989: Holocene paleohydrology of central Alberta: testing the general-circulation-model climate simulations; Canadian Journal of
Earth Sciences, v. 26, p. 1826-1833.
Servant-Vildary, S. and Roux, M.
1990: Multivariate analysis of diatoms and water chemistry in Bolivian
saline lakes; Hydrobiologia, v. 197, p. 267-290.
Smol, J.P., Curnming, B.F., Douglas, M.S.V., and Pienitz, R.
1995: Inferring past climatic changes in Canada using paleolimnological
techniques; Geoscience Canada, v. 21, p. 113-1 18.
Smol, J.P., Walker, I.R., and Leavitt, P.R.
1991: Paleolimnology and hindcasting climatic trends; Verhandlungen
der Internationalen Vereinigung von Limnologen, v. 24,
p. 1240-1246.
Stockton, C.W. and Meko, D.M.
1983: Drought recurrence in the Great Plains as reconstructed from
long-term tree-ring records; Journal of Climate and Applied
Meteorology, v. 22, p. 17-29.
Street-Perrott, F.A. and Roberts, N.
1983: Fluctuations in closed lakes as an indicator of past atmospheric circulation patterns; in Variation in the Global Water Budget, (ed.)
F.A. Street-Perrott, M. Berman, and R. Ratcliffe; D. Reidel
Publishing Co., Dordrecht, The Netherlands, p. 331-345.
ter Braak, C.J.F.
1986: Canonical correspondence analysis: a new eigenvector technique
for multivariate direct gradient analysis; Ecology, v. 67,
p. 1167-1178.
1988: CANOCO - A FORTRAN program for canonical community ordination by [partial] [detrended] [canonical] correspondence analysis, principal components analysis, redundancy analysis (version
2.1); Agricultural Mathematics Group, Wageningen, The
Netherlands, Technical Report LWA-88-02,95 p.
1990: Update Notes: CANOCO version 3.10; Agricultural Mathematics
Group, Wageningen, The Netherlands, 35 p.
ter Braak, C.J.F. and van Dam, H.
1989: InferringpH from diatoms: a comparison of old and new calibration
methods; Hydrobiologia, v. 178, p. 209-223.
Turner, L.J.
1994: 210Pbdating of lacustrine sediments from Cleanvater Lake (Core
064, Station CWSI), Saskatchewan; National Water Research
Institute, Burlington, Ontario, Contribution 94-141,27 p.

Vance, R.E. and Last, W.M.


1994: Paleolimnology and global change on the southern Canadian
Prairies; in Current Research 1994-B; Geological Survey of
Canada, p. 49-58.
1996: Stop 4: Oro Lake; in Landscapes of the Palliser Triangle, (ed.)
D.S. Lemmen; Guidebook for the Canadian Geo~norphology
Research Group Field Trip, Canadian Association of Geographers
1996 Annual Meeting, Saskatoon, Saskatchewan, p. 26-27.
Vance, R.E., Clague, J.J., and Mathewes, R.W.
1993: Holocene paleohydrology of a hypersaline lake in southeastern
Alberta; Journal of Paleolimnology, v. 8, p. 103-120.
Vance, R.E., Last, W.M., and Smith, A.J.
1997: Hydrologic and climatic implications of a multidisciplinary study
of late Holocene sediment from Kenosee Lake, southeastern
Saskatchewan, Canada; Journal of Paleolimnology, v. 17,
p. 365-393.
Vance, R.E., Mathewes, R.W., and Clague, J.J.
1992: 7000 year record of lake-level change on the northern Great Plains:
a high resolution proxy ofpast climate; Geology,v. 20, p. 879-882.
Van Stempvoort, D.R., Edwards, T.W.D., Evans M.S.,
and Last, W.M.
1993: Paleohydrology and paleoclimate records in a saline prairie lake
core: mineral, isotope and organic indicators; Journal of
Paleolimnology, v. 8, p. 135-147.
Wilson, S.E.
1996: Paleosalinity reconstructions using diatoms from lakes in western
Canada; Ph.D. thesis, Queen's University, Kingston, Ontario,
187 p.
Wilson, S.E., Cumrning, B.F., and Smol, J.P.
1994: Diatom-salinity relationships in 111 lakes from the Interior Plateau
of British Columbia, Canada: the development of diatom-based
models for paleosalinity reconstructions; Journal of
Paleolimnology, v. 12, p. 197-221.
1996: Assessing the reliability of salinity inference models from diatom
assemblages: an examination of a 219 lake dataset from western
North America; Canadian Journal of Fisheries and Aquatic
Sciences, v. 53, p. 1580-1594.
Wilson, S.E., Smol, J.P., and Sauchyn, D.J.
1997: A Holocene paleosalinity diatom record from southwestern
Saskatchewan, Canada: Harris Lake revisited; Journal of
Paleolimnology, v. 17, p. 23-31.
Xia, J., Haskell, B.J., Engstrom, D.R., and Ito, E.
1997: Holocene climate reconstructions from tandem trace-element and
stable-isotope composition of ostracodes from Coldwater Lake,
North Dakota, U.S.A.; Journal of Paleolimnology,v. 17,p. 85-100.
Yansa, C.H.
1998: Holocene paleovegetation and paleohydrology of a prairie pothole
in southern Saskatchewan, Canada; Journal of Paleolimnology,
v. 18, p. 429-441.

Groundwater inputs to a closed-basin saline lake,


Chappice Lake, Alberta
S.J. irks' and V.H. ~ e m e n d a ~
Birks, S.J. and Remenda, V.H., 1999: Groundwater inputs to a closed-basin saline lake, Chappice
Lake, Alberta; in Holocene Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate Change on the Southern Canadian
Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological Survey of Canada, Bulletin 534,
p. 81-93.

Abstract: The chemistry and dynamics of groundwater in the Chappice Lake basin were monitored to
delineate potential groundwater sources. Slug tests revealed that Chappice Lake is surrounded by a high
hydraulic-conductivity system ( ~ = 1 0 rn.s-l),
-~
with the possibility of a confining low hydraulic-conductivity (K=10-7 m-s-l)layer at depth. Measured hydraulic heads and water-table elevations show
strong annual fluctuations that correspond with seasonal changes in recharge. The chemical compositions
of groundwater springs entering the lake resemble the water chemistry of bedrock aquifers (Na-HC03) and
surficial aquifers (Mg-Ca-Na-HC03), suggesting that both shallow and deep groundwater systems recharge
the lake. However, a strong horizontal component to flow and small upward vertical hydraulic gradients are
suggestive of discharge by a shallow groundwater system. Groundwaters sampled at elevations below the
lake resemble lake water in both their chemical (Na-SO4) and isotopic compositions, suggesting that they
originated as outflow from the lake. This indicates that the lake is not hydrologically closed, despite having
no surface outflow.

RCsum6 : La chimie et la dynamique des eaux souterraines dans le bassin

du lac Chappice ont Ct6


Ctudites afin de d6limiter des sources d'eau souterraines potentielles. Des essais de puits ont montr6 que le
lac Chappice est entour6 d'un systbme ?
haute
i
conductivitt hydraulique ( ~ = 1 0m.s-I);
- ~ il est possible qu'il
existe en profondeur une couche impermkable de faible conductivitt hydraulique (K=10m7rns-l). Les
mesures de la charge hydraulique et de la hauteur de la nappe phrtatique tCmoignent de fortes fluctuations
annuelles qui correspondent aux changements saisonniers de recharge. La composition chimique des
sources d'eaux souterraines se d6versant dans le lac resemble 2 celle des eaux des aquifhres du substratum
rocheux (Na-HC03) et des aquifkres de surface (Mg-Ca-Na-HC03), ce qui laisse supposer que des
systkmes d'eaux souterraines tant superficiels que profonds rechargent le lac. Toutefois, la prtsence dans
17Ccoulementd'une forte composante horizontale et de faibles gradients hydrauliques verticaux ascendants
semble indiquer qu'il y a Bcoulement h partir d'un systkme d'eaux souterraines peu profond. Les eaux
souterraines Cchantillonn6es h divers niveaux sous le lac ressemblent aux eaux du lac par leur composition
chimique (Na-SO4) et isotopique, ce qui sugghre qu'elles proviennent d'un d6bit sortant issu du lac. Cela
indique que le lac n'est pas hydrologiquement fermC, en dtpit du fait qu'il est dtpourvu de debit sortant en
surface.

Department of Earth Sciences, University of Waterloo, Waterloo, Ontario N2L 3G1


Department of Geological Sciences and Geological Engineering, Queen's University, Kingston, Ontario K7L 3N6

GSC Bulletin 534

INTRODUCTION
The fossil and sedimentary composition of cores from
closed-basin groundwater-fed lakes has been used as a source
of high-resolution paleoenvironmental data for the Great
Plains. Although these data are used to reconstruct
paleohydrological conditions, little is generally known about
the hydrogeology of these lakes and how it may influence
their sensitivity to changes in climate (Remenda and Birks,
1999). In lakes where groundwater inflow is significant, it is
possible for groundwater to dampen the connection between
lake levels and climate (Street-Perrott and Harrison, 1985).
At Chappice Lake in southeastern Alberta (Fig. I), mineralogical and paleobotanical indicators of past lake levels
(Vance et al., 1992; Vance et al., 1993) and stable isotopes in
cellulose and carbonate minerals (Padden, 1996) have been
used to reconstruct aspects of Holocene climate and
paleohydrology. These studies have implicitly assumed that
Chappice Lake is fed by a shallow aquifer that is sensitive to
climate change. Such inflow has been deemed critical to
allow Chappice Lake to exist, without significant surface
inflow, in an area where potential evaporation greatly
exceeds precipitation. In follow-up to these studies, a detailed
hydrogeological study of Chappice Lake was undertaken, in
which physical and chemical characteristics were used to
identify sources of groundwater to the lake (Birks, 1995;
Birks and Remenda, 1999). This paper presents some results
of this study to illustrate features of groundwater flow to topographically closed basin lakes.

BACKGROUND
Lakes on the Great Plains display a wide range of brine concentrations and compositions, from dilute Ca-HC03 to
hypersaline Na-SO4 (Last, 1992; Donovan and Rose, 1994;
Last, 1999). It has been suggested that the sources of solutes
to these lakes may include dissolution of deep Paleozoic
evaporites by upward-flowing water (Grossman, 1968),
evaporation of groundwater from shallow Cretaceous and

Tertiary bedrock sources (Wallick, 1981), and evaporation of


groundwater from shallow surficial aquifers (Witkind, 1952;
bonovan and Rose, 1994; see ~ast,1999).The extreme range
in water compositions has been explained by a combination
of varying source-water compositions (Donovan and Rose,
1994) and different groundwater flux ratios (Wood and Sanford, 1990).These parameters are unique to each lake and can
therefore vary over small areas.
The typical prairie groundwater profile consists of a
nested series of shallow, intermediate, and deep groundwater
systems, each with different physical and chemical characteristics (see Remenda and Birks, 1999). The regional slope,
topography, and position of the lake relative to the major discharge area determine the type of groundwater system feeding the lake (Winter, 1977; Almendinger, 1990). Lakes
situated in the lower part of a drainage basin, near a regional
discharge area such as a river, are more likely to receive flow
from deep, regional groundwater systems. Lakes located
higher in the regional setting, farther away from major outlets, are more likely fed by shallower groundwater systems.
Groundwater flow in the vicinity of lakes has been modelled and studied in the field (Winter, 1976; Shaw and Prepas,
1990). Vertical-section, steady-state numerical simulations
indicate that the height of the water table, the position and
hydraulic conductivity of surrounding aquifers, and lake
depth strongly influence the position and shape of the flow
field (Winter, 1976). In some hydrogeological settings, vertical variation in the flow direction results in loss of groundwater from the lake to the underlying flow system (Winter,
1976). This advection of lake water outward from below the
lake is one means by which lakes that appear to be
hydrologically closed can lose water and solutes. Diffusion
and density-driven flow are other mechanisms by which
closed-basin lakes can lose solutes (Wood and Sanford,
1990).If the densities of saline lake brines exceed fresh-water
hydraulic heads in the basin, a hydrodynamically unstable situation may be created, allowing fingers of brine to sink into
the underlying aquifers (Wood and Sanford, 1990; Ferguson
et al., 1992).

Aspen parkland
Bunchgrass steppe

Figure 1.

Location and bathymetry of Chappice Lake. The


lake has no surjhce outlets, but was fed by the
stream entering the northeast corner (see Fig. 2)
until a weir was constructed in I976. Bathymetry
based on an August 1989 survey.

S.J.Birks and V.H. Rernenda

The response of lake levels to changes in climate depends


on both the sensitivity to changes in the surface water balance
(precipitation and evaporation) and the climatic sensitivity of
groundwater inputs. Sensitivity to precipitation and evaporation is dependent on groundwater-surface water interactions
(Donovan and Rose, 1994), which are related to the surface
area of the lake and the permeability of the surrounding aquifer (Almendinger, 1990). Shallow lakes with large surface
areas, situated on low-permeability deposits, are likely to
have lake levels that are most sensitive to changes in the
amounts of precipitation and evaporation. The sensitivity of
groundwater inputs relates to the type of groundwater discharging into the lake. Lakes situated in the discharge area of
a deep, regional groundwater system would likely have
groundwater inputs that are relatively constant and insensitive to climate. In contrast, lakes situated in the discharge area
of a shallow, local system would have groundwater inputs
that are more sensitive to changes in climate.

STUDY AREA
Chappice Lake (lat. 5010'N, long. 11022'W, elev. 721 m
a.s.1.) is situated in a meltwater channel on the edge of a large,
internally drained, upland area approximately 15 km from the

South Saskatchewan River (Fig. 1). The channel, with local


relief of about 30 m, cuts through areas of low-relief ground
moraine to the north and south, and hummocky moraine to the
west (Fig. 2; Berg and McPherson, 1972). Runoff is concentrated in small closed basins that form important recharge
points on hummocky moraine (Meyboom, 1967;Fortin et al.,
1991), but result in disintegrated, poorly developed surface
drainage.

Geology
Most of the Chappice Lake area is underlain by the Upper
Cretaceous Oldman Formation (Fig. 3), a freshwater sandstone and sandy shale unit with ironstone bands and bentonite
(Stewart, 1940). It has an estimated well yield of
20-95 L-rnin-l and is the aquifer in which most water wells
are developed (Stevenson and Borneuf, 1977). The overlying
Bearpaw Formation, a dark grey marine shale that subcrops
upland areas, has an estimated well yield of less than
5 ~arnin-l(Stewart, 1940; Stevenson and Borneuf, 1977).The
Foremost Formation, which underlies the Oldman, consists
of an upper carbonaceous shale unit with an estimated well
yield of less than 5 ~emin-l,and a lower sandstone unit with a
yield estimated at 20-95 ~ . m i n - (Stevenson
l
and Borneuf,
1977).

Figure 2. Airphoto mosaic showing Chappice Lake (1947) and surrounding area. A series of
meltwater channels (mwc) converge at Chappice Lake, with the main channel continuing eastward.
HM, hummocky moraine; GM, ground moraine. NAPLA217.55-158, -159, -161, -162.

GSC Bulletin 534

Stevenson and Borneuf (1977) indicated that Chappice


Lake is situated over the thalweg of a buried bedrock valley
(Fig. 3). On the southern Prairies, these valleys, which may be
up to 5 km wide, were formed by pre-Quaternary to
mid-Wisconsinan rivers and subsequently infilled by
glaciogenic deposits, eliminating almost all evidence of them
from the surface topography (Stalker, 1969; Klassen, 1989).
Insufficient data on depth to bedrock are available from the
vicinity of Chappice ~ a k to
e determine the exact location or
dimensions of the buried valley. Depending on the nature of
the fill between the bedrock surface and the lake, the buried
valley could be a significant feature in controlling lake
hydrology. If filled with permeable sand and gravel, buried
valleys have the potential to serve as aquifers (van der Kamp,
1986; Shaw and Prepas, 1990). However, a thick clay-rich fill
(e.g. till) would reduce the potential of the buried valley to act
as a groundwater source for the lake. A correlation between
the location of buried glacial valleys and the occurrence of
salt lakes has been noted (Witkind, 1952).
Surficial deposits consist primarily of till in upland areas
and glaciofluvial sediments within the meltwater channels
(Berg and McPherson, 1972). The channel occupied by
Chappice Lake is one of several subparallel, southeast-trending meltwater channels that converge in the vicinity of the lake. These channels are generally filled with
greater than 20 m of sand and gravel.

Water well samples

Climate
Chappice Lake is located in a mid-latitude steppe, with a
potential evapotranspiration of 567 rnm.a-l, exceeding the
mean annual precipitation of 363 inm.a-' (Stevenson and
Borneuf, 1977). Most of the precipitation is received in the
late spring, but high evaporation results in little infiltration or
runoff. Melting of snow accumulated during the winter is
likely the greatest source of recharge to the system, as has
been documented elsewhere on the Prairies (Meyboom,
1967; Keller and van der Kamp, 1988).

Hydrology
Chappice Lake is located in a topographically closed basin
with no surface outlets. The surface-water drainage basin has
an area of about 156 km2, whereas the surface area of the lake
is only 2.1 km2.The ratio of drainage basin area to lake area
suggests a large potential area available for overland runoff
(Winter, 1977). Since the upland areas are composed of
fine-grained material with a low permeability and the valley
deposits are mainly sand and gravel, the valleys likely form
significant conduits for groundwater flow from runoff generated on upland areas.
Prior to installation of a weir in 1976, Chappice Lake was
fed by a stream entering the northeast corner (Fig. 1). The former stream bed now supports a large marsh, suggesting that
subsurface flow along that path is still significant. Comparison of stream channel areas occupied by marshes and standing water upstream and downstream of the weir site obtained

Piezometer samples

I
Estimated well
yield (~.min-')

Quaternary
Bearpaw
Oldman

r1

Foremost

Undifferentiated
sand, silt, diamicton

Dominantly
shale

Dominantly
r lsandstone

Figure 3. Schematic cross-section of the Chappice Lake basin with estimated well
yields for bedrock aquifers (Stevenson and Borneuj 1977). Depth of buried bedrock
channel depicted below the lake and nature of the Quaternary fill are not known.
Private water wells in the vicinity of Chappice Lake were sampled to determine the
geochemical signatures of bedrock aquifers (G
),,
and shallow surficial aquifers
(Gshnllow), which were then compared with the geo~%emistry
of groundwater sampled
from piezometers installed in the basin (Gnbove,
Gbelow, and G,e,,surfnce) to determine
the type of groundwater entering the lake. Reference datum for 'above', 'below', and
'near surface' is August 1994 lake level.

S.J. Birks and V.H. Remenda

from airphotos flown prior to (August 1947) and after installation (August 1985) suggests that continued high groundwater inputs along the stream channel have minimized the
weir's effect on the annual water budget. The groundwater
component includes flow to the lake through the
high-permeability valley deposits in the meltwater channel.
Airphotos of Chappice Lake taken between 1947 and
1985 also document large fluctuations in lake surface area
during this interval (Vance et al., 1993). Although the lake
presently consists of two isolated basins (Fig. I), water levels
in the past have been high enough to breach the divide
between the basins. Changes in lake level, as measured from
airphotos, have been used with precipitation, lake evaporation, and bathymetry data to evaluate the historical climatic
sensitivity of lake levels and groundwater inputs (Birks arid
Remenda, 1999). Changes in lake volume can be calculated
using equation 1 (Street-Pesrott and Harrison, 1985).
where:
AV = change in volume of the lake, measured from airphotos
(m3>

AL = area of the lake (m2)

PL = precipitation on the lake (m)


EL = evaporation from the lake (m)
R = surface runoff from the catchment to the lake (m3)

Table I. Summary of water budget analyses for

Chappice Lake (from Birks and Remenda, 1999).


Volume changes (AV) were calculated for each interval
covered by airphotos. Total inputs to the lake (@) were
greatest when precipitation deficits were highest (most
negative).

Time
interval

Volume
change AV
(ma)

Preclpltation deflclt
A,(P,-E,)
(m3)

(m3.a-')

Total Inputs

cP
(m3)

(m3-a-')

1947-1 952

126000

-3100000

-622000

3230000

1952-1 955

270000

-1 900000

-625000

2150000

647000
715000

1955-1 962

-249 000

-6 200 000

-887 000

5 960 000

852 000

1962-1970

74 900

-6 800 000

-851 000

6 890 000

861 000

1970-1 976

-78 500

-5 400 000

-896 000

5 300 000

883 000

1976-1 980

-120 000

-3 400 000

-849 000

3 280 000

819 000

1980-1 981

-92 000

-81 000

-810 000

718 000

718 000

1981-1985

-46 000

-3 000 000

-760 000

3 000 000

749 000

deficit or rate of input. Time lags and feedbacks between lake


levels and groundwater inputs may account for the lack of
correlation for individual time periods (Birks and Remenda,
1999). The largest groundwater inputs to the lake occurred
when the precipitation deficit was high and the lake level was
low. Increased removal of surface water seems to have
resulted in increased groundwater inputs. Thus, lowering of
the lake level during surface water removal results in higher
hydraulic gradients and increased groundwater flux (Birks
and Remenda, 1999).

D = surface discharge from the lake (m3)


Gi = groundwater inputs to the lake (m3)
Go = groundwater outflow from the lake (m3)
Surface discharge (D) equals zero because there are no
surface outlets. The runoff and groundwater flux were
lumped in a total input term ( a ) , simplifying to equation 2
where:
= total input to the lake from

surface sources and groundwater


Lake area varied from a minimum of 1.49km2 in 1985 to a
maximum of 2.21 km2 in 1955, during the 38 years covered
by airphotos. Changes in lake volume over this period were
small (generally <5%) compared to changes in the precipitation deficit (AL (PL- EL) (Table 1). To maintain such a small
range in volume change over periods with pronounced
changes in precipitation deficits requires a total groundwater
input almost equivalent to the precipitation deficit. At times,
increases in the precipitation deficit were associated with
decreases in lake level, a response consistent with assumptions made in the paleoclimate reconstruction (Vance et al.,
1993). However there were also times when the assumed
relationships between lake level and climate did not apply:
for example, lake volume increased between 1962 and 1970
despite continued high precipitation deficits (Table 1).
Results of this simple water budget indicate that variations in
lake volume are not a simple function of either precipitation

FIELD METHODS AND ANALYSES


Identification of the sources of inflowing groundwater to
Chappice Lake was based upon comparison of the physical
and chemical characteristics of groundwater entering the lake
with those of shallow (Gshallow)
and deep (Gdee ) groundwater systems in the vicinity of the lake, as weyl as those
reported in the literature (Fig. 3). Groundwater flow entering
the lake was monitored using four standpipe piezometers,
four multilevel piezometers, a continuous water-table monitor, and 21 minipiezometers installed during the 1993 and
1994field seasons (Fig. 4,5, Table 2). Daily measurements of
depth to static water level in the piezometers were made
during August 1994 and monthly measurements continued
until August 1995. Piezometers CH 93-02, CH 93-03, and
CH 94-01 were installed at sites expected to be removed from
lake effects, and were used with water samples taken from
private water wells in the area to characterize shallow
groundwater chemistry. Private water wells screened in bedrock units were sampled to characterize the chemistry of deep
groundwater. Slug tests were performed on 11 of the wells,
and hydraulic conductivities were determined using the
Hvorslev method (Freeze and Cherry, 1979).
Standpipe and multilevel piezometers were installed with
a hollow-stem auger drill rig. Prior to installation, a testhole
was drilled with a solid stem auger, with sediment samples
being collected from the auger at about 1.5 m intervals.
Standpipe piezometers consisted of a single well constructed

GSC Bulletin 534

Table 2. Piezometer installation elevation and hydraulic conductivities


calculated from slug test data, Chappice Lake (from Birks and
Remenda, 1999).
Elevation
of top of

ground
surface

Hydraulic
conductivity,
K (m.s-' )

0.61

CH 94-01

3.7 x 10.'

732.10

3.1 x 10"

2.8x 10"

CH 94-026

CH 94-02C

1.3 x 10"

1.7~
10'~
4.4 x 10-6
-

4.9

104

1.7 x 1 v 5
6.8 x 10''

2.9 x 10-5
-

1.61
~0 . ~

out of polyvinylchloride (PVC) pipe. Multilevel piezometers


consisted of a bundle of piezometers constructed from a centre stalk of PVC pipe with polyethylene (PET) tubing
attached. Each multilevel piezometer bundle was assigned a
label (e.g. CH 94-02), while individual piezometers with
intake screens at different depths within the bundle were
labelled using letters, with A corresponding to the deepest
piezometer (Fig. 4).
Minipiezometers were constructed from PET tubing with
a 0.13 m screen and installed to a depth of 0.9 m along lines
radiating from the lake edge (Fig. 5). Each rninipiezometer
line was given a label (e.g. CH 94-08), while individual
minipiezometers were identified by their position in the line
using lower case roman numerals increasing outward from
the lake edge (Fig. 5). Minipiezometer lines were placed in
areas of groundwater springs, as identified by the presence of
salt flats and salt-intolerant vegetation. The water-table well,
consisting of a tube of steel casing slotted below the water
table, was installed to a depth of 10 m using a hydraulic rotary
drilling rig. A continuous, type-F water-level monitor was set
up at the water-table well on November 1, 1994.
Private water wells were selected for sampling on the
basis of their proximity to Chappice Lake and the availability
of documentation on the water-supplying units, using Alberta
Environment's well-registry database. Wells screened in single bedrock units were preferred; however, where wells of
this type were not present, wells screened over more than one
unit were used. Electrical conductivity (EC), pH, and temperature of groundwater samples were measured in the field.
Samples were run through a 0.45 pm filter and titrated for
alkalinity within one day of collection. Chemical analyses
were performed by the inorganic chemistry laboratory at the
Geological Survey of Canada in Calgary, Alberta. Samples

were analyzed for fluoride, chloride, bromide, nitrate, phosphate, sulphate, lithium, sodium, ammonium, potassium,
magnesium, and calcium using liquid chromatography.
Three lake-water samples, three groundwater samples,
and a surface-water sample from the vicinity of the groundwater springs were analyzed for 2~ and 1 8 0 at the Environmental Isotope Laboratory, University of Waterloo,
Waterloo, Ontario. Results are given in per mil (%o)notation
relative to Vienna Standard mean ocean water (VSMOW)
and calculated using the formula:
%O= ((Rsa,v[e - RvS&fOw) RVSMOw)
where R = 180/160or 2 ~ 1 ~ .

luou

RESULTS
Cross-sections prepared from testhole logs showed that the
surficial deposits in the meltwater channel consist primarily
of poorly sorted sand (Fig. 4), with the lothpercentile ranging
from 0.02 to 0.008 cm (Deyell, 1995). The piezometer intake
for CH 94-01 was installed in silt at a depth of 20 m, beneath
massive diamicton (interpreted as till) that forms the steep
north slope of the meltwater channel. It is not known whether
the silt occurs as a lens within the till or underlies the till. The
lake is bounded on the south by a terrace within the meltwater
channel composed mainly of sand (Fig. 4). Piezometer CH
93-02 was located on this terrace. The total thickness of sand
surrounding the lake within the meltwater channel is not
known. Although piezometer CH 94-03A, located on the
exposed divide between the two lake basins, was installed
20 m below ground surface in silty clay, it is not known
whether this marks the base of the glaciofluvial fill or is
simply a clay lens within the fill.

S.J. Birks and V.H. Remenda

CH 94-02, -03, -04, and -05 had the same general trend as
CH 93-02 and -03, albeit with greater variation. At CH 94-02
and -03, many small fluctuations in elevation were recorded
in addition to the annual cycle. An annual cycle in hydraulic-head elevations was also observed at CH 94-01; at this
site, however, seasonal fluctuations were of smaller amplitude and were delayed at least nine months, with the maximum value occurring in January.

During August 1994, the lake level was surveyed at


721.0 m a.s.1. Hydraulic heads measured in all piezometers
were greater than the lake surface elevation, indicating that
Chappice Lake presently lies in a closed hydraulic-head contour. Fluctuations in hydraulic-head elevations at CH 93-02
and -03 were synchronous (Fig. 6), despite being 2.5 km
apart. Maximum values occurred during spring and minimum
values at the end of the summer. Fluctuations in the
water-table elevation and hydraulic-head elevations at

.-0

...

740

CH 94-05
e.
..

...

%....\
'-

-,-

710

700

a. 45
5 .z;>;:i--:!?
----. .1,730
......:
. :.. . . . . .-. . ....'......:.::?-.-..
:,,,; :,

. . . . .. . ..: : .. . ... . . ..
. ... .. ... .. . ..
. . .
,

)Sand

Till

-.:'::::;.:::

---,-,-,

~~Io::--:D.

....................
,. .. .
. , , . . , , 2550:,--:, c,
... . .... ... .. .. . . .". ... .. . ". ... . . 8300'.:-::~
. . .............
. ,. ,

)Lacustrine - clay

11 850 A ' A '

Electrical
conductivity
(pS.cm-')

1p,iCpmeter

_ -,Estimated
extent

Figure 4. North-south and east-west cross-sections showing electrical conductivity


(EC) values (in bold). Groundwaters from CH 93-02, CH 93-03, and CH 94-01 are
relatively dilute and appear unafected by Chappice Lake. The thickness of the
lacustrine deposit is a minimum estimate based on length of sediment cores recovered
by vibracoring (Vance et al., 1993).

Figure 5.
Locations of minipiezometers at Chappice Lake,
installed at 0.9 m depth along lines chosen to
intercept areas of groundwater spring discharge, as indicated by salt-intolerant vegetation. High electrical conductivities of
groundwaterfi-om near the lake edge are due to
mixing with lake water. Salt flats occur where
the water table is near the s u ~ a c eLocations
.
of
some standpipe and multilevel piezometers
shown for reference.

Minipiezorneler line
Minipiezometer

Electrlcal
conductivity (vS,cm.')

L!%iU

GSC Bulletin 534

v Lake water

* Shallow water (Gshalloiv)


well samples

v Deep water (Gdeep)


well samples

CH 93-02 A CH 83-03-0- Water table

Figure 6. Fluctuations in hydraulic heads at CH 93-02 and


-03, Chappice Lake, measured over a two-year period,
compared with water-table elevations. Sites are located
approximately 2.5 km apart.

% of total millieauivalents
Cations
per litre
Anions

Figure 7. Piper plot of groundwater samples taken from


private water wells in shallow surjicial (Gsh,lOw, <61 m
depth) and bedrock (Gdeep, 61-122 m depth) aquifers,
compared to Chappice Lake water. Lake-water samples were
collected between 1989 and 1995.

Vertical hydraulic gradients (i) were calculated between


the highest and lowest intakes for each of the three multilevel
piezometer bundles, CH 94-02/03, -04, and -05, using equation 3 (Maathuis and van der Kamp, 1994):
Piezometer samples
taken above (G,-)
lake level

where
-04 8 0 5

Piezometer samples
taken below (G,,,,)
lake level

Ap, = hydrostatic pressure difference


pa, = average fluid density of the formation between piezometers

Az = elevation difference between screen midpoints of piezometer


p = pore pressure in monitoring well
Density-corrected hydraulic gradients show the potential
for upward flow at CH 94-05 (i = 0.03) and CH 94-02/03 (i =
0.0009), and downward flow at CH 94-04 (i = -0.0005). The
water chemistry data required to make the density corrections
were only available for August 1994. Consequently, the gradients presented here represent only a snapshot of groundwater flow at that time. Horizontal hydraulic gradients were
greatest in minipiezometer lines to the northeast (CH 94-10)
and south of the lake (CH 94-08), suggesting a strong horizontal component to flow in these areas.
The lake water is hypersaline (extremely high EC,
Table 3, Fig. 4) and has a Na-SO4 composition (Fig. 7). Calcium and bicarbonate concentrations are low (Table 3), suggesting removal by precipitation. This is supported by
mineralogical analyses of the core that showed high percentages of calcium and magnesium carbonates (Vance et al.,
1993).
Groundwater obtained from private water wells located
outside of Chappice Lake basin plots in two distinct groups
on a Piper diagram (Fig. 7).Wells screened at depths less than
61 m in shallow surficial aquifers have a Mg-Ca-Na-HCO,

% % of total rniil!equivalents T?
Cations

per l~tre

Anions

Figure 8. Piper plot of piezometer groundwater samples


from depths below (Gbelow)and above (Gabove)the elevation
of Chappice Lake. Comparison with Figure 7 shows that CH
94-01 and CH 93-02plot in the same area as shallow surficial
aquifers, while CH 93-03 plots in the same area as bedrock
waters.

composition and ECs less than 1500 p ~ . c m - lWells


.
screened
at depths greater than 61 m in bedrock aquifers have a
Na-HC03 composition (Fig. 7) and higher ECs (16901920 yS-cm-l).
Groundwater sampled within the Chappice Lake basin
exhibits a wide range of concentrations (Fig. 4,5) and compositions (Fig. 8,9), and can be grouped on the basis of sampling
location. Samples from piezometers screened below lake

S.J. Birks and V.H. Remenda


Table 3. Water chemistry and isotopic analyses, Chappice Lake (from Birks and Remenda, 1999).
ANIONS
Sample

Chloride
(PP~)

Bromide
(PP~)

CATIONS
Llthlum
(PP~)

Sulphate
(PP~)

Sodium
(PP~)

Potasslum
(PP~)

Magnesium
(PP~)

Calclum
(PP~)

pH

Alkallnlty
(meq)

Electrical
conductivity
(~S.cm")

Isotopes
S'80
( )

6'H

-8.56

-89.79

-17.61

-145.30

(L)

Private water wells

Shallow wells screened in shallowaquifsrs< 61 m In depth


PW1

69

n.d.

276.9

0.96

145.0

PW2

42.2

n.d.

223.6

0.99

105.2

PW3

12

n.d.

13.6

n.d.

10.5

PW4

47.1

n.d.

263.7

0.96

136.1

PW5

34.2

n.d.

547.7

0.97

103.4

Deep wells screenedin bedrock squMers behveen 61 and 121 m In depfh


PW6

71.2

33.5

184.6

0.95

PW7

35.9

n.d.

313.6

1.09

504.5
560.3

PW8

39.4

n.d.

436.9

1.01

556.5

PW9

78.3

21.0

137.6

1.02

851.6

PWlO

49.6

n.d.

556.6

1.00

595.8

PW11

58.8

42.1

470.9

1.01

482.5

PW12

54.9

n.d.

697.4

1.09

581.6

CH 93-02

5.3

n.d.

153.4

0.47

92.3

CH 93.02

33.8

n.d.

156.3

n.d.

69.0

CH 93.03

41.9

n.d.

335.1

0.15

346.5

Plezometer samples

CH 94-01

56.9

766.3

0.22

218.8

1193.5

55.6

14031.6

0.96

6059.1

CH 94-028

694.2

52.3

11976.6

0.69

4771.0

CH94.03A

1536.3

59.8

16383.2

1.34

6658.6

CH 94.038

1104.5

n.d.

15238.6

1.22

6390.6

CH 94.03C

1052.4

n.d.

14637.9

0.92

6687.5

CH 94-04A

234.5

36.1

3003.9

1.07

1596.1

CH 94-048

173.5

43.3

2301.0

0.23

1099.1

CH 94-04C

146

2.6

2023.2

0.24

1014.9

CH 94-04D

185

43.8

2198.2

0.24

1166.5

CH 94-02A

n.d.

CH 94-05A

1.8

0.0

6.0

0.43

3138.1

CH 94-058

244.4

4.6

5115.3

0.26

2256.1

n.d.

1238.3

0.12

668.3

1765.6

0.20

835.8

671.3

0.16

350.1

CH 94-05C

64.5

CH 94-05D

165.8

CH 94-05E

56.1

43.2
n.d.

Mlnlplezometer samples
CH 94-081

0.5

0.1

11.7

0.43

2669.4

CH 94-06ii

496.7

6.2

4630.9

0.36

2507.4
196.0

CH 94-08111

19.3

n.d.

356.3

0.15

CH 94.061~

39.9

n.d.

175.4

0.15

92.2

CH 94.091

1987.9

66.3

52321.6

1.67

23507.7

CH 94-0911

2638.7

76.9

18709.4

1.65

9916.9

CH 94-09111

2194.2

69.9

17429.6

1.47

9276.6

CH 94-091v

2291.6

76.1

23151.0

1.84

11766.5

CH 94-09v

244.1

44.8

3121.5

0.33

1651.4

CH 94.101

339.6

5.7

4128.8

0.32

1850.9

CH 94-1011

67.4

1.2

1109.6

0.20

743.0

CH 94-10111

33.9

n.d.

592.0

0.19

529.8

CH 94-10iv

70.9

1.3

1094.1

0.16

764.1

CH 94.111

47.4

n.d.

746.2

0.20

617.6

CH 94.121

36.7

n.d.

752.9

0.19

800.8

CH 94.12111

26.5

n.d.

877.7

0.19

588.8

CH 94-131

64

1.6

1309.3

0.22

527.9

CH 94-13li

146.7

2.7

2325.3

0.24

1464.9

CH 94.141

372.8

6.3

2886.9

0.31

1829.7

32.2

n.d.

404.4

0.16

504.5

CH 94-1411
Lake water
August. 1992

23000

n.a.

93500.0

16.7

50900.0

~ugust,1993

9540

n.a.

106000.0

10.5

50930.0

August, 1994

20500

n.a.

72700.0

21.7

34200.0

January, 1995

41000

n.a.

65600.0

30.3

44100.0

May, 1995
Surface water (sample taken In the viclnlty of a groundwater spring)
January, 1995
n.8. - not analyzed
n.d. - not detected

434

n.a.

4470.0

0.3

2300.0

40.0

14.6

76.7

7.9

673.0

10100

GSC Bulletin 534

level (CH 94-02, -03, -04, and -05) have high ECs (Fig. 4) and
plot in a similar area on a Piper diagram as lake water
(Na-SO4 water). Groundwater from piezometers screened
above lake level (CH 93-02 and -03) have low ECs and are
interpreted as being isolated from the influence of the lake.
Groundwater from CH 94-01, screened in the silt unit at the
base of the clay, also has a fairly low EC and will be included
with this group, since its location within or below the till
serves to effectively isolate it from the lake. The composition
of groundwater sampled from CH 93-02 and CH 94-01 is
similar to groundwaters from shallow surficial aquifers and
plots in a similar area on the Piper diagram (Fig. 8). Groundwater from CH 93-03 has a Na-HC03 composition resembling groundwaters sampled from bedrock aquifers.
Near-surface groundwater also exhibits a large range in
concentration (Fig. 5 ) and composition (Fig. 9). Electrical
conductivities range from 900 to more than 50 000 y ~ . c m - l ,
with the highest concentrations from piezometers located
along the lake shore or in salt flat areas. Samples from these
areas have compositions similar to that of lake water (Fig. 9).
The more dilute groundwater samples were obtained from
discharging groundwater springs. These samples plot in two
groups on a Piper diagram: the first similar to bedrock waters
and the second similar to shallow surficial aquifers (Fig. 9).
The extremely high ECs (Fig. 5) and compositional sirnilarity to lake water (Fig. 9) suggest that minipiezometers on
the lake edge are sampling lake water present in pore fluids.
Salt flat areas occur where the water table intercepts the
ground surface. The high ECs of near-surface groundwater
sampled on salt flats reflects evaporative concentration of
groundwater at the surface.
The 6180 and 6 2 of~lake water and groundwater spring
samples, as well as high-salinity groundwater sampled below
lake level at CH 94-05A, are plotted relative to a meteoric
water line for the Prairies (PMWL, 6 2 = ~8.9 x 6180 + 21.1;
Fig. 10). Preferential removal of the light isotopes of water

% % of total millieouivalents
Cations

per lit&

Near-surface samples:
lake shore (G ~ , , f l a lahe
~ shore)
and salt flats @nea,u&m,
oats)
Near-surface samples:
springs (Gnea,,,fla,,
ti,,^^)

Anions

Figure 9. Piper plot of minipiezometer groundwater samples


from the edge of Chappice Lake fGneap,,flafnce, lakeshore)!
salt flats fGneur-sLlrf,ce,
salt
and from SPrlngs
(Gnear-su,face, springs).

pm

Lake water
CH 94-05A
A Surface
sample,
spring

Figure 10. 6 1 8 0 and 6 2 of~ water sampled from


groundwater springs (CH 94-08iii and -1Oiii)and Chappice
Lake, plus high-EC water sampled at depth below lake level
(CH 94-05A), relative to a prairie meteoric water line
(PMWL, h2H = 8.9 x 6180 + 21.1) and a local evaporation
line (LEL, h2H = 5.6 x 6180 - 47.7). Continuation of LEL
back to the PMWL should indicate the starting isotopic
composition of the water. Groundwaterfrom the springs was
sampled from pooled sug5ace water (open triangle) and from
rninipiezometers (solid triangles).
during evaporation will result in a shift away from the
PMWL. As a result, lake-water samples are enriched in the
heavy isotopes relative to precipitation.

DISCUSSION
Chappice Lake is situated in a sand-dominated, high hydraulic-conductivity groundwater system bounded to the north by
the low hydraulic-conductivity till. Assuming that the lacustrine sediments described by Vance et al. (1992) are continuous across the lake bottom, groundwater entry is likely
restricted to the nearshore zone of the lake. This is consistent
with both modelling and field studies that have shown
groundwater inputs are usually highest along the lakeshore
and decrease towards the centre of the lake (Winter, 1976;
Almendinger, 1990; Shaw and Prepas, 1990). Since
Chappice Lake lies within a closed hydraulic-head contour,
groundwater is directed towards the lake from all directions.
However, the rate of groundwater entry along the north shore
is likely much lower than the rest of the lake, because it is bordered by low-permeability till. The scarcity of groundwater
springs along the north shore of the lake supports this
assertion.
Calculation of hydraulic gradients was based upon the
average water density between sampling points, an approach
that may have introduced error to the calculations. It should
also be emphasized that these gradients represent only a snapshot of groundwater flow in the vicinity of the lake, and do not
depict the likely seasonal fluctuations in the flow field.

S.J. Birks and V.H.

Remenda

Table 4. Summary of water chemistry analyses, Chappice Lake.

Chemical facies refers to the field on the Piper diagram in which the
majority of samples plot.
Electrical

Sampling location

facies

Gdeep

(Wallick, 1981)

n.a.

Gnear sudace, lake shore, and


ssll flats

Lake

3510-50 OoO
10400-50000

~a-SO,
Na-SO4

(Wallick, 1981)
n.a.

n.a.

n.a.

-1.19to -10.79

-59.03to -106.5

n.a.- not available


Nonetheless, these data do suggest that groundwater gradients in the Chappice Lake basin are near neutral and not spatially pervasive. The weak vertical hydraulic gradients,
indicating both downward and upward flow of water in
piezometers sulrounding the lake, suggest discharge from a
shallow groundwater system, and are inconsistent with discharge from an intermediate or deep groundwater system. In
a regional discharge area, discharge from a deep groundwater
system results in strong vertical hydraulic gradients and the
upward flow of water throughout the basin.

Piezometers that are not expected to reflect outflow from the


lake (Gab& are CH 93-02 and -03, located above lake level,
and CH 94-01, which is isolated from the lake by till. Samples
from these sites have greater potential to preserve a geochemical
signature indicative of their source. While CH 93-02 has a
Ca-Na-Mg-HCOg type water, typical of shallow surficial sandy
aquifers in the region, CH 93-03 has a Na-HC03 composition
similar to that of bedrock waters. Piezorneter CH 94-01 has a
composition typical of groundwater from tills on the southern
Prairies.

The electrical conductivities and chemical facies identified as characteristic of bedrock aquifers and shallow
surficial aquifers (Table 4) agree with those reported in the
literature (Wallick, 1981; Donovan and Rose, 1994). However, differentiating on the basis of geochemical signature
between shallow and deep groundwater systems as the original source of water entering the lake is complicated by the
effect of the lake on the surrounding groundwater.

As was the case with the piezometer samples, the geochemical composition of groundwater springs not influenced
by lake water or evaporation resembles both bedrock waters
(e.g. Na-HC03 type water at CH 94-10iii) and shallow
surficial waters (e.g. Na-Mg-HC03 type water at CH 94-8iii).
The 6180 and 6 2 of~the groundwater springs are similar and
plot on the PMWL, which is consistent with a shallow
groundwater system. Extension of the LEL from lake water
compositions to the PMWL provides an indication of the isotopic composition of the source water to the lake prior to
evaporation. While the isotopic composition of the groundwater springs plots at the intersection of the PMWL and the
LEL, it also lies within the range measured from Belly River
Formation waters in east-central Alberta (Wallick, 1981).
Hence the isotopic composition of bedrock water is not sufficiently different from that of meteoric water to allow
unequivocal determination of the source of groundwater to
the springs.

The high-salinity groundwaters sampled below lake level


(Gbelow,from piezometers CH 94-02, -03, -04, and -05) have
concentrations and compositions consistent with outflow of
hypersaline water from the lake (Table 4). These waters are
far more concentrated (i.e. EC = 11 850 p~.cm-l)and have a
different composition than groundwater sampled from deep
bedrock aquifers, suggesting that they are not discharge from
a deep groundwater system. The enriched isotopic signal
found at CH 94-05A (6180 = -12.85%0,6 2 =~-1 16.5%0)provides further evidence that the high-salinity groundwater
sampled at elevations below the lake is outflow from the lake.
Local precipitation and groundwater derived from local precipitation will plot along the PMWL (Fig. 10). The local
evaporation line (LEL) at Chappice Lake, determined from
the isotopic composition of surface waters, is 6 2 =~5.6 x
6180 - 47.7 (Figure 10). The isotopic signature of
groundwater from CH 94-05A plots on the LEL between
spring water and lake water, and is likely the result of mixing
of evaporatively enriched hypersaline lake waters with shallow groundwater during outflow from the lake.

The variability in local geochemical and isotopic groundwater signatures, as well as the climate sensitivities of different groundwater sources, must be considered in paleoclimate
and paleohydrological reconstructions based on Chappice
Lake sediments. Increased rates of groundwater input during
periods of low lake levels have been identified in the historical hydrological record of Chappice Lake, likely driven by
increased horizontal hydraulic gradients (Birks and
Remenda, 1999). Such gradient changes would likely be
greatest in the high-permeability sand aquifer surrounding
the lake, which in turn would change the ratio of the different
groundwater sources and ultimately affect the chemical and

GSC Bulletin 534

isotopic composition of the lake water. Hence reconstructed


changes in lake chemistry may relate to changes in
source-water composition rather than evaporation.
The presence of a high-salinity groundwater system
below Chappice Lake demonstrates that the basin is not
hydrologically closed. The thick deposits of
low-permeability lacustrine deposits underlying the lake suggest that diffusion is the dominant process currently operating. Early in the lake's history, prior to deposition of the
fine-grained lacustrine sediments, the sandy aquifer would
have formed the base of the lake. At this time, the hydraulic
conductivity of the basal sand unit may have been high
enough for density-driven advection to occur. Fingers of
dense saline lake water could have descended from the lake,
forming the saline system intercepted by piezometers
screened below lake level.
The loss of water from lakes has important implications
for brine evolution and the mineralogy of lake sediments.
Since the groundwater flux ratio (the ratio of groundwater
outflow from the lake to groundwater inflow to the lake) ultimately determines how evaporation will affect brine evolution and mineral precipitation, leakage from the lake could
potentially result in the maintenance of steady-state solute
conditions under continued evaporation (Wood and Sanford,
1990).

SUMMARY
Results of this hydrogeological investigation illustrate the
complexity of the groundwater regime of Chappice Lake.
The high salinity, geochemistry, and isotopic signature of
groundwater sampled below the lake show that the lake is not
hydrologically closed. The chemical composition of groundwater springs entering the lake resembles both bedrock aquifers ( N a - H C 0 3 ) and shallow surficial aquifers
(Mg-Ca-Na-HC03), suggesting contributions from both
shallow and deep groundwater systems. However, the presence of a large horizontal component to groundwater flow,
small upward vertical hydraulic gradients, and fluctuations in
hydraulic head corresponding to seasonal changes in
recharge are all consistent with the discharge of a shallow
groundwater system. The majority of available evidence
therefore suggests that Chappice Lake is currently sustained
by the discharge of a shallow groundwater system, a conclusion that supports the assumption of Vance et al. (1993) in
their ~aleoclimaticreconstructions. This conclusion is also
supported by previous geochemical simulations of brine evolution, which demonstrated that groundwater with a chemistry similar to bedrock aquifers could not have been the
dominant source of water over the entire history of the lake
(Birks and Remenda, 1999).

ACKNOWLEDGMENTS
We thank D.S. Lernrnen and R.E. Vance for assistance during
this project. We also wish to acknowledge the Caven family,
the Pensose family, and the City of Medicine Hat for allowing

access to their property. Chilako Drilling of Lethbridge and


Joden Drilling of Medicine Hat are thanked for their skilled
and efficient help. All of the groundwater chemistry data used
in this study was kindly provided by H.J. Abercrombie of the
Geological Survey of Canada, Calgary. We thank
S. Roadhouse and J. Wei for field assistance, and B. Gorham
for technical assistance. The manuscript benefited from
reviews by W.M. Last and G. van der Kamp. This study was
supported by a Palliser Triangle Global Change Research
Grant from the Geological Survey of Canada, a Natural
Sciences and Engineering Research Council of Canada
(NSERC) research grant to V.H.R., and a Queen's University
graduate research fellowship to S.J.B.

REFERENCES
Almendinger, J.E.
1990: Groundwater control of closed basin lake levels under steady-state
conditions; Journal of Hydrology, v. 112, p. 298-318.
Berg, T.E. and McPherson, R.A.
1972: Surficial geology, Medicine Hat, Alberta, NTS 72L; Alberta
Research Council, Map 142, scale 1:250 000.
Birks, S.J.
1995: Hydrogeological investigation of Chappice Lake, southeastern
Alberta: groundwater inputs to a saline basin; M.Sc. thesis, Queen's
University, Kingston, Ontario, 208 p.
Birks, S.J. and Remenda, V.H.
1999: Hydrogeological investigation of Chappice Lake, southeastern
Alberta: groundwater inputs to a saline basin; Journal of
Paleolimnology, v. 21, p. 235-255.
Deyell, C.
of llydraulic conductivity: a review of
1995: Laboratory measurcme~~ts
laboratory methodology using results from Chappice Lake, Alberta;
B.Sc. thesis, Queen's University, Kingston, Ontario, 184 p.
Donovan, J.J. and Rose, A.W.
1994: Geochemical evolution of lacustrine brines from variable-scale
groundwater circulation; Journal of Hydrology, v. 154, p. 35-62.
Ferguson, J., Jacobson, G., Evans, W., White, I., Wooding, A.,
Barnes, C.J., and Tyler, S.
1992: Advection and diffusion of groundwater brines in modern and
ancient salt lakes, Nulla groundwater discharge complex, Murray
Basin, southeast Australia; irz Water-Rock Interaction:Proceedings
of the 7Ih International Symposium on Water-Rock Interaction,
(ed.) Y.K. Kharaka and A.S. Maest; A.A. Balkema, Rotterdam,The
Netherlands, p. 643-647.
Fortin, G., van der Kamp, G., and Cherry, J.A.
1991: Hydrogeology and hydrochemistry of an aquifer-aquitard system
within glacial deposits, Saskatchewan, Canada; Journal of
Hydrology, v. 126, p. 265-292.
Freeze, R.A. and Cherry, J.A.
1979: Groundwater;Rentice-Hd Inc., Englewood Cliffs, New Jersey, 604 p.
Grossman, I.G.
1968: Origin of the sodium sulphate deposits of the northern Great Plains
of Canada and the United States; United States Geological Survey,
Professional Paper 600-B, p. B 104-B 109.
Keller, C.K. and van der Kamp, G .
1988: Hydrogeology of two Saskatchewantills, II.Occurrenceof sulfate and
implicationsfor soil salinity;Journal of Hydrology, v. 101,p. 123-144.
Klassen, R.W.
1989: Quaternary geology of the southern Canadian Interior Plains; in
Chapter 2 of Quaternary Geology of Canada and Greenland, (4.)
R.J. Fulton; Geological Survey of Canada, Geology of Canada,
no. 1, p. 138-173. (also Geological Society of America, The
Geology of North America, v. K-I).
Last, W.M.
1992: Chemical composition of saline and subsaline lakes of the northern
Great Plains, western Canada; International Journal of Salt Lake
Research, v. 1, p. 47-76.

S.J. Birks and V.H. Remenda


Last, W.M. (cont.)
1999: Geolimnology of the Great Plains of western Canada; in Holocene
Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Maathuis, H. and van der Kamp, G.
1994: Subsurface brine migration at potash waste disposal sites in
Saskatchewan: final report; Saskatchewan Research Council,
Publication No. R-1220-10-E-94,90 p.
Meyboom, P.
1967: Mass-transfer studies to determine the groundwater regime of permanent lakes in hummocky moraine of western Canada; Journal of
Hydrology, v. 5, p. 117-142.
Padden, M.C.
1996: Holocene paleohydrology of the Palliser Triangle from isotope
studies of lake sediments; M.Sc. thesis, University of Waterloo,
Waterloo, Ontario, 108 p.
Remenda, V.H. and Birks, S.J.
1999: Groundwater in the Palliser Triangle: an overview of its vulnerability and potential to archive climate information; in Holocene
Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Shaw, R.D. and Prepas, E.E.
1990: Groundwater-lakeinteractions,11. Nearshore seepage patterns and
the contribution of groundwater to lakes in central Alberta; Journal
of Hydrology, v. 119, p. 121-136.
Stalker, A.MacS.
1969: Quaternary stratigraphy in southern Alberta, report Ik sections near
Medicine Hat; Geological Survey of Canada, Paper 69-26.28 p.

Stevenson, D.R. and Borneuf, D.M.


1977: Hydrogeology of theMedicine Hat area, Alberta; Alberta Research
Council, Report 75-2, 11 p.
Stewart, J.S.
1940: Geology of Redcliffe, Alberta; Geological Survey of Canada, Paper
41-1 1 (map only, scale 1:253 440).
Street-Perrott, F.A. and Harrison, S.P.
1985: Lake levels and climate reconstruction; in Paleoclimate Analysis
and Modelling, (ed.) A.D. Hecht; Wiley, New York, p. 291-340.
Vance, R.E., Clague, J.J., and Mathewes, R.W.
1993: Holocene paleohydrology of a hypersaline lake in southeastern
Alberta; Journal of Paleolirnnology,v. 8, p. 103-120.
Vance, R.E., Mathewes, R.W., and Clague, J.J.
1992: 7000 year record of lake-level change on the northern Great Plains:
a high-resolution proxy of past climate; Geology, v. 20, p. 879-882.
van der Kamp, G.
1986: The groundwater resources of Saskatchewan's deep buried-valley
aquifers; in Proceedings of the 1986 Canadian Hydrology Symposium. No. 16, June 3-6, Regina, Saskatchewan; National Research
Council of Canada, p. 529-543.
Wallick, E.I.
1981: Chemical evolution of groundwaterin a drainage basin of Holocene
age, east-central Alberta, Canada; Journal of Hydrology, v. 54,
p. 245-283.
Winter, T.C.
1976: Numerical simulation analysis of the interaction of lakes and
groundwater; United States Geological Survey, Professional Paper
1001,45 p.
1977: ~lassific&onof the hydrologic settings of lakes in the north-central
United States; Water Resources Research, v. 13, p. 753-767.
Witkind, I.J.
1952: The localization of sodium sulfatedeposits in northeasternMontana
and northwestern North Dakota; American Journal of Science,
v. 250, p. 667-676.
Wood, W.W. and Sanford, W.E.
1990: Ground-watercontrol of evaporite deposition; Economic Geology,
V.85, p. 1226-1235.

Mineralogy, lithostratigraphy, and inferred


geochemical history of North Ingebrigt lake,
Saskatchewan
Yuqiang shangl and William M. ~ a s t '
Shang, Y. and last, W.M., 1999: Mineralogy, lithostratigraphy, and inferred geochemical history
of North Ingebrigt lake, Saskatchewan; in Holocene Climate and Environmental Change in the
Palliser Triangle: A Geoscientific Contextfor Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.)D.S. Lemmen and R.E. Vance; Geological Survey of Canada,
Bulletin 534, p. 95-1 10.

Abstract: North Ingebrigt lake, a small, hypersaline playa basin in the Great Sand Hills of southwestern
Saskatchewan, contains one of Canada's thickest Holocene terrestrial salt sequences. The 10 000-year long
lacustrine sequence from the basin consists of well indurated salt, with only minor mud and organic debris.
Cores lack obvious bedding, colour variation, or other visible sedimentary structures. The mineral suite
consists mainly of hydrated Na-, Mg-, Ca-, and (Na+Mg)-sulphates, -carbonates, and -chlorides. This long
and apparently uninterrupted sequence of soluble salts implies that the lake was characterized by high salinity throughout its Holocene history. Preliminary chronostratigraphy indicates that accumulation rates have
varied greatly in the past. Similarly, variations in salt mineralogy suggest that the basin experienced significant changes in brine composition. These evaporites permit reconstruction of specific ion activities and
ratios in the brine through time, providing the first detailed evidence of the complexity of brine evolution in
a closed-basin lacustrine environment in western Canada.

R6sum6 : Le lac North Ingebrigt, petit bassin playa hypersalin dans les dunes Great Sand Hills du
sud-ouest de la Saskatchewan, contient une des plus Cpaisses sCquences de sels continentaux holoci?nes au
Canada. La sequence lacustre du bassin, longue de 10 000 ans, se compose de sel trks indurC renfermant de
faibles quantitCs de boue et de dCbris organiques. Les carottes ne prCsentent ni stratification manifeste, ni
variation de couleur, ni autre structure ~Cdimentairevisible. La suite minerale est composCe essentiellement
de sulfates, de carbonates et de chlorures hydratCs de Na, Mg, Ca et Na+Mg. L'existence de cette longue
sCquence apparemment ininterrornpue de sels solubles permet de supposer que le lac se caractkrisait par une
forte salinitt pendant I'ensemble de son histoire holoc&ne.Les Ctudes chronostratigraphiques priliminaires
indiquent que les taux d'accumulation ont fortement variC par le passC. De msme, les variations de la
minkralogie des sels laissent penser que le bassin a subi des changements importants quant 2 la composition
des saumures. Ces Bvaporites permettent la reconstitution dans le temps des activitCs et des rapports
ioniques spCcifiques de la saumure, fournissant ainsi les premikres indications dCtaillCes sur la complexit6
de 1'Cvolution des saumures dans un milieu lacustre de bassin fermC dans I'Ouest du Canada.

Department of Geological Sciences, University of Manitoba, Winnipeg, Manitoba R3T 2N2

GSC Bulletin 534

INTRODUCTION
Mineralogical and lithostratigraphic investigations of lake
sediments have traditionally provided paleolimnologists with
an important but usually nonquantitative understanding of
such factors as water depth, drainage-basin weathering conditions, tectonic setting, and sediment provenance
(Gierlowski-Kordesch and Kelts, 1994; Dearing, 1997). In
contrast, paleobiological parameters have often been called
upon to contribute a somewhat more quantitative perception
of the lake. This is still true in many lacustrine settings: biological indicators such as ostracodes, chrysophytes, diatoms,
chironomids, and plant macrofossils can provide exceptional
quantitative information on past water salinity, major-ion
chemistry, nutrient level, temperature, and depth in most
fresh and marginally saline lakes (e.g. Berglund, 1986;Davis,
1990; Sullivan and Charles, 1994).
However, in large areas of western North America as well
as many other regions of the world, saline to hypersaline surface waters dominate the landscape. Climatic and hydrological conditions during much of the mid- to early Holocene in
western Canada gave rise to greatly elevated water salinities,
even in lakes where modern ionic concentrations are low.
These high-salinity brines are often accompanied by strongly
alkaline conditions which curtail the use of siliceous
microfossils (Evans, 1993; Gel1 et al., 1994). Shallow water
and associated high-energy oxidizing conditions at the sediment-water interface further limit the potential of
paleobiological work by creating poor preservation conditions for uollen and macrofossils. At the same time, these
saline and hypersaline lacustrine settings can provide investigators with an extraordinarily explicit record of past changes
in water chemistry through the endogenic and authigenic
minerals preserved in the sediment. As emphasized by Teller
and Last (1990), the evaporites are particularly useful
because their mere presence (versus abundance) can provide
unambiguous data on the thermodynamic conditions of the
precipitating solution.
Evaporite formation and related physicochemical conditions have been subjects of long-standing sedimentological
interest and study over the last two centuries (Eugster, 1971;
Multhauf, 1978). Numerous chemical models have been
established to predict the sequence of salts precipitated
during evaporative concentration of a brine. In general, two
approaches have been used to determine the stable mineral
assemblages and solution composition: direct evaporation
experiments and thermodynamic and solubility calculations.
Since the experimental work of Van't Hoff and his successors
in the early part of this century, a large amount of data has
accumulated on the solubility of salts in different aqueous
systems (Assarsson, 1950a, b, c; Stewart, 1963; Braitsch,
1971;Eugster, 1971).These data have been successfully used
to graphically predict the sequence of major minerals precipitated from evaporating brines by means of concentration-temperature diagrams and phase (or Jfinecke) diagrams
(Harvie et al., 1982). This graphical approach is best applied
to relatively simple binary and common-ion aqueous systems, and cannot readily manage the more complex ionic systems that commonly occur in natural brines (Hardie, 1984).

To solve this problem, several generations of computer models based on equilibrium thermodynamics have emerged over
the past two decades. Although many of these models are limited to dilute aqueous solutions with ionic strengths less than
0.7, some are valid to high ionic strengths (Harvie et al., 1984;
Weare, 1987).
Experimental and theoretical geochemical models provide a powerful tool that can assist in deciphering past chemical and hydrological conditions and depositional processes
during evaporite formation in a sedimentary basin. In this
paper, we use these mineral solubility models as the theoretical basis for interpreting the observed evaporite mineral
assemblages and chemical compositions of brines in North
Ingebrigt lake, Saskatchewan, and the hydrological conditions under which these evaporites were deposited. Although
commonly applied to brine reconstruction in marine and lacustrine settings elsewhere, this was the first such effort to
apply these models to lake deposits of the Great Plains of
North America.

METHODS
This paper is based on mineralogical and lithostratigraphic
analyses of a single 23 m long sediment core retrieved from
the North Ingebrigt lake basin in February 1994. As most of
the sedimentary fill in the North Ingebrigt basin consists of
highly indurated salt, conventional paleolirnnological coring
techniques (e.g. Livingstone piston coring, Vibracoring, hollow-stem Shelby-tube augering) were not able to penetrate
andlor recover the sediment. A specially modified, air-cooled
diamond-drillinglcoring technique was used to retrieve core.
Details regarding this drilling equipment and unique coring
technique, not previously used in Quaternary paleolirnnology
in Canada, are found in Shang (1999). The diamond drilling
afforded excellent recovery of large-diameter (6 cm), undisturbed sediment cores. Because a compressed air (rather than
water) circulation system was used, the highly soluble salts
experienced little chemical modification or dissolution
during retrieval. In addition to the 23 m core reported in this
paper, cores and auger samples were also collected from
other sites in North Ingebrigt lake. Analytical results from
these other sections and intrabasinal correlations are presented elsewhere (Shang, 1999).
The core was extruded in the field, cut into sections
approximately 2 m long, wrapped in plastic, and transported
to the University of Manitoba, where it was stored in a cold
room (4C) until sampled. In the laboratory, core sections
were cut longitudinally using a rock saw, photographed, and
one of the halves was subsampled at 2 cm intervals for various
physical, chemical, and mineralogical analyses. Subsamples
for 14cdating and other analyses were taken at irregular intervals. Subsampling was done in a cold room and subsamples
were stored at 4 O C to minimize postcollection modification
of the mineral suite.
Mineralogical analyses were conducted using X-ray diffraction (XRD) techniques (Last, 1996; Shang, 1999). A
nonaqueous medium was used in sample preparation, and
subsamples were stored in a temperature- and humidity-

Y. Shang and W.M. Last

controlled environment until analyzed. Mineral identification was aided by the use of an automated search-match computer program (Marquart, 1986). Percentages of the various
minerals were estimated from the bulk mineral
diffractograms using the intensity of the strongest peak for
each mineral, as outlined by Schultz (1964) and Last (1980).
Nonstoichiometry of the dolomite and calcite was determined by examining the displacement of the dIo4peak on a
detailed (slow) XRD scan (Goldsmith and Graf, 1958) and
calculated according to Hardy and Tucker (1988). All
detailed evaporite and carbonate mineralogy was deciphered
from detailed (slow) XRD scans. Moisture content, organic
matter content, and total carbonate mineral content were
evaluated by weight loss on heating to temperatures of 45"C,
500C, and 1000C, respectively (Dean, 1974). Duplicate
samples were prepared and analyzed for both mineralogy and
weight loss every 50 cm. These replicate analyses indicate
that precision of the data reported here is approximately 1%
for weight loss and 7% for mineralogy.

(-46" to 45C) and precipitation variations. Fox Valley,


20 km north of the basin, receives approximately 300 mm of
precipitation annually, and has a mean daily temperature of
-14OC in January and +21C in July.
Salts have been commercially extracted from numerous
saline wetlands and playas in the vicinity of North Ingebrigt
lake for more than 70 years (Tomkins, 1953, 1954; Last and
Slezak, 1987). One of North America's largest sodium sulphate mines and processing plants is located a few kilometres
south at Ingebrigt Lake. Salts have also been extracted from
Chain (local name), Snakehole, Corral, Boot (local name),
Verlo (local name), and Vincent lakes, all within 30 km of the
North Ingebrigt basin. Although North Ingebrigt has, to date,
not been mined, the lake is presently under mineral lease, and
Cole (1926) identified a total anhydrous Na2S04 reserve of
about 3 x lo6 t.

Various statistical analyses were performed on the mineralogical and lithostratigraphic data in order to evaluate
parameter interrelationships (Pearson product-moment linear correlation analysis and R-mode factor analysis), and to
define stratigraphic units (cluster analysis). All statistical
evaluation was done using the computer program SAS (SAS
Institute Inc., 1989). Details regarding these statistical procedures are provided in Sheskin (1997) and Shang (1999).
Modified versions of the computer programs WATEQF
(Truesdell and Jones, 1974; Rollins, 1989), MIX2 (Plummer
et al., 1975), and PHRQPITZ (Plummer et al., 1988; Shang,
1999) were used to help determine the thermodynamic activity and molar ratios of the the-mica1 species.

MODERN NORTH INGEBRIGT LAKE


General setting
North Ingebrigt lake (lat. 5024.5'N, long. 109"22.5'W, elev.
695 m a.s.1.) occupies a narrow riverine basin in the Great
Sand Hills of southwestern Saskatchewan (Fig. I), about 100
km west of Swift Current. The basin probably originated
about 14 000-13 000 BP as a glacial meltwater spillway
superposed on the buried preglacial Johnsborough Valley
(David, 1964; Christiansen, 1979). The area is underlain by
up to 60 m of till and glaciofluvial sediments, which in turn
overlie about 3 krn of Phanerozoic sedimentary rocks. The
approximately 300 km2 watershed of the lake is characterized
by poorly integrated drainage and hummocky to gently rolling topography consisting mainly of sand flats, undulating
moraine, and stabilized sand dunes (Epp and Townley-Smith,
1980; Saskatchewan Research Council, 1985).
The closed basin of North Ingebrigt lake is about 3 km
long and 750 m wide, with a high-stage surface area of
1.25 km2. The lake is located in the most arid portion of the
northern Great Plains. The Sand Hills area experiences a cold
continental climate characterized by extreme temperature

Figure 1. Generalized distribution of modern sedimentary


facies and location of springs in North lngebrigt lake. Inset
shows location of the lake within the northern Great Plains of
western Canada.

GSC Bulletin 534

Hydrogeology and hydrological setting

Modern sedimentology

North Ingebrigt lake is one of a series of distinct, separate,


riverine salt playas occupying this 50 km long, north-trending
meltwater channel. Closure of the North Ingebrigt basin is
small and a water level of more than approximately 1.5 m
above the present-day surface will spill over to Ingebrigt
Lake, less than 1 km to the south. The spill point for the entire
Ingebrigt-North Ingebrigt chain of lake basins is approximately 20 m above the present-day surface of the valley floor.

North Ingebrigt lake is a salt-dominated playa. Except for a


narrow, nearshore fringe of organic-rich, fine-grained and
poorly sorted clastic sediments, the modern lake sediments
are almost entirely very soluble evaporitic salts (Fig. 1).
These endogenic precipitates of the modern salt-pan sedimentary facies comprise mainly hydrated Na- and
Mg-sulphates, including mirabilite (Na2S04.10H20),
bloedite (Na2Mg(S04)2.4H20), epsomite (MgS0,.7H20),
and leonhardtite (MgS04.4H20). Gypsum (CaS04.2H20),
halite (NaCl), despujolsite (Ca3Mn(S04)2(0H)6.3H20),and
protodolomite (disordered, nonstoichiometric MgCa(C03)2)
are commonly present in small to trace amounts in the upper
few centimetres. The inorganic fraction of the detrital sediments of the mud flat, sand flat, and surrounding colluvium
facies are composed of clay minerals (mostly illite and
expandable lattice clays), feldspars, carbonate minerals
(mainly dolomite), quartz, and ferromagnesian minerals.

The modern hydrology and hydrochemistry of North


Ingebrigt lake and surrounding drainage basin were discussed by Shang (1999), although David (1964), Ruffel
(1968), and Rutherford (1967, 1970) also provided important
regional and local groundwater and hydrochemical data.
Typical of most of the shallow, salt-dominated playas in
western Canada, the hydrological budget of North Ingebrigt
lake is dominated by groundwater inflow and evaporation.
There is no perennial streamflow into this topographically
closed basin, and evaporation exceeds mean annual precipitation by a factor of 3 (Canadian National Committee for the
International Hydrologic Decade, 1978). Groundwater enters
the lake via both diffuse inflow and several discrete springs.
The lake exhibits typical playa characteristics, filling with
water during the spring and early summer and usually drying
completely by late summer or early fall. Maximum water
depth is generally less than 50 cm.
North Ingebrigt lake water is usually hypersaline, with
salinities in the range 100-300 g . ~ - total
l
dissolved solids
(TDS). Although strong seasonal variation occurs in brine
chemistry, the water is alkaline (pH > 8), stron ly dominated
by Na+ and S042-, and nearly depleted of Ca and HC03(Fig. 2). Inflowing groundwater is relatively dilute (less than
5 g . ~ -TDS),
l
with higher proportions of Ca2+and HC03-.

i?+

Modern sedimentation in North Ingebrigt lake is controlled largely by the seasonal (or periodic) sequence of
flooding, evaporative concentration, and desiccation of the
playa surface. Like many other saline playas in the region,
sedimentary processes are dominated by: 1) formation of salt
crusts, efflorescent crusts, spring deposits, and intrasedimentary salts; 2) subaqueous cumulate and bottom salt
precipitation; 3) physical reworking and redistribution of
clastic salt and accretionary salt grains; 4) formation of salt
cements; 5 ) irregular dissolution of surface salt crusts and formation of solution pits on the playa surface and within the
near-surface sediment; 6) formation of karst chimneys; and
7) mud diapirism and subsequent reworking of fine-grained
siliciclastic material. These processes were discussed in
detail by Last (1984, 1989a) and Shang (1999). The salt

Figure 2. Mean water composition (milli-equivalent %) of North Ingebrigt lake (A), inflowing springs (+),
and groundwater from near-sutface Cretaceous bedirock (v)and glacial deposits (.).

Y. Shang and W.M. Last

America (e.g. Smith and Friedman, 1986; Last, 1989b) indicate that the evaporite mineral suite is largely stabilized
within a few tens of centimetres below the playa surface.

playas in the Ingebrigt chain contain several of the largest


Holocene saline karst features identified in the northern Great
Plains (Last, 1993a).
The paragenesis of the mineral suite in North Ingebrigt
lake is complex. In a groundwater-dominated, hypersaline
salt pan/playa environment such as that found in today's
basin, the distinction between endogenic and authigenic rninera1 components is obscure (see also Sonnenfeld, 1984;
Smoot and Lowenstein, 1991). Salts that may have formed as
true endogenic precipitates (i.e. derived from within the water
column of the lake) are very likely dissolved and reprecipitated numerous times at the sediment-water interface as
the composition of the overlying brine changes during the
year, or at the surface of the playa upon subaerial exposure
and desiccation. Likewise, both fresh and saline pore waters
within the upper few centimetres of the playa surface can
readily modify the precipitated salts. Although seasonal and
longer term mineralogical changes and accompanying diagenesis of the nondetrital suite in North Ingebrigt sediment
are complex processes, observations from numerous other
saline playas in western Canada and other parts of North

Organic
matter

STRATIGRAPHY
General description
The stratigraphic sequence recovered from North Ingebrigt
lake consists almost entirely of massive, coarsely crystalline,
well indurated salt. The sediment of the upper few metres of
the section is somewhat more friable, with slightly finer crystal size, than deeper parts of the core. Minor amounts of
fine-grained mud occur sporadically throughout the section.
Organic matter content is low, averaging less than 2% (Fig. 3;
Table 1). The salt is generally translucent but ranges from
opaque to clear. The sediment varies from colourless to white
(N8; MunsellTMColour Chart), grey (N4 to N7), pale yellow
(5Y8/4), light olive grey (5Y6/2), and pale red (7.5R612).
Except for subtle changes in clarity, colour, and degree of

Total
carbonate

Allogenic vs.
(endogenic plus
authigenic)

Mineral
facies

Missing core

k LkEMissin core

~ m

~ ~

Missing core

mm

F..******
~~

LFL
...m..mm....

10

20

Yo

.*.*...*..

m.m..mmm.m..

30

10

20

Yo

30

25

50
%

Figure 3. Stratigraphic variation in percentages of organic matter, total carbonate


content, and proportion of total inorganic allogenic material versus endogenic+
authigenic components in North Ingebrigt lake core. Heavy horizontal dotted lines
delineate mineral facies (see Table I ) . Gaps in plots represent sections in which there
is missing core.

GSC Bulletin 534

Table 1. Summary of average characteristics of mineral facies from North lngebrigt


lake core.
Mineral facies
1

Na-SO,
facies

Mg-C03Na-SO4
facies

Ca-SO4Na-SO4
facies

Muddy
Na-SO4
facies

Total clay minerals (%)


Quartz (%)
Total feldspar minerals (%)
Potassic feldspars (%)
Plagiociase (%)
Total calcite (%)
Low-Mg calcite (%)
Magnesian calcite (%)
MgC03 in Mg-calcite (mol%)
Total dolomite (%)
Stoichiometric dolomite (%)
Protodolomite (%)
CaCOSin protodolomite (mol%)
Magnesite (%)
Aragonite (%)
Ca suiphates (%)
Na sulphates (%)
Mg sulphates (X)
Chlorides (%)
Pyrite (%)
Allogenic fraction (%)
(Endogenic+authigenic) fraction (%)
Total oraanlc matter (%)

6.7
0.6
1.8
0.1
1.7
0.1
0.1
0.1
8.7
1.5
0.3
1.2
62.6
1
0.3
3.4
83.9
0.4
0.1
0.1
10.7
89.3
1.3

12.2
1.1
0.5
0.1
0.4
<0.1
<0.1
0.1
6.3
1.9
0.5
1.3
63.8
1
0.2
17.2
65.5
0.1
<0.1
n.d.
15.8
84.2
2.5

9.1
2.5
11.7
0.4
11.3
0.1
0.1
0.1
9.4
1.2
0.5
1.6
62.5
0.3
0.1
5.5
72
n.d.
<0.1
n.d.
24.6
75.4
1.9

5
Muddy
Mg-calciteNa-SO4
facies

Dolomitic
Na-SO,
facies

Core
average

9.9
3.5
3
1.4
1.5
0.7
0.1
0.6
12.2
3.3
1.7
1.5
60.3
0.1
0.1
6.8
76.2
0.2
<o. 1
n.d.
20.4
79.6
2.4

5.2
2.6
3.1
0.4
2.7
0.2
<0.1
0.2
10.4
3.1
0.8
2.3
60.9
0.3
0.1
5.1
79.5
0.7
<0.1
<0.1
11.6
88.2
2.1

n.c. not calculated


n.d. - not detected

Table 2. Endogenic and authigenic mineralogy of North lngebrigt lake core.


Mineral name
~ompositlon
Carbonate mlnerals
Aragonite
CaC03

Occurrence'
Common

Artinite
Ankerite
Benstonite
Calcite
Gaylussite
Kutnohorite
Magnesite2
Magnesian calcite
Minrecordite
Protodolomite
Zemkorite
Sulphate mlnerals

Mg,C03(OH),.3H,0
C~(F~,MS)(CO,)~
Ca7~a6(C03)13
CaCO,
Na,Ca(C03),5H20
Ca(Mn,Mg)(CO&
M9c03
(Mg$a,.JCO,
CaZn(COS),
(Mg,Ca,.,)(C0,)2
Na2Ca(C03)2

Very rare
Rare
Very rare
Rare
Very rare
Very rare
Common
Common
Very rare
Very common
Very rare

Arcanite
Bloedite
Despujolsite
Epsomite
Eugsterite
Gargeyite
Gypsum

(K,NH,),SO,
Na&lg(S0,),.4H20
C+Mn(SO,),(OH).3H,O
MgS0,.7Hz0
Na,Ca(S0,),.2H20
K,Ca,(S04)6.H20
CaS04.2H,0

Rare
Very common
Rare
Common
Very rare
Very rare
Very common

Mlnerai name
Composltlon
Sulphate mlnerals (cont.)
Polyhalite
&CazMg(S0,),.2H,0

Occurrence'

Potasslum alum
Leonhardtite
Thenardite

KAi(SO,),.l2H2O
MgSO4.4H,O
Na2S0,

Very rare
Very rare

Bonshtedlte
Bradleyite
Burkeite

NhFe(P0,)(C03)
Na,Mg(PO,)(COJ
Na,(SO,)CO,

Very rare
Very rare
Very rare

Rare

Very common
Watteviilelte
Na2Ca(S0,),.4H,0
Very rare
Carbonate-sulphate, carbonate-sulphate-chorde and
carb~nate-phosphateminerals

Hanksite
KNa,(CO,),(SO,),CI
Rapidcreekite
C%(COJSO4.4H,O
Tychite
Na6Mg,SO4(CO3
Chloride minerals
Bischofite
MgCI2.6H,O
Carnallite
KMgCI,,.GH,O
Dougiasite
KFeC14.2H,0
Halite
NaCl
Nltrate and borate mlnerals
inderborite
CaMgB,O,,.H,O
Niter
KNO,
Soda Niter
NaNOs
Nitrobarite
Ba(NO,),
Other minerals

Common
Very rare
Very rare
Very rare
Very rare
Very rare
Rare

Rare
Very rare
Hexahydrite
MgS0,.4H20
Common
Very rare
Kieserite
MgS0,-H,O
Rare
Rare
Krausite
Fe,(S0,),.2H20
Very rare
Leightonite
K2Ca,C~(S0,),.2H,O
Very rare
Mercaiiite
KHSO,
Very rare
Pyrite
FeS,
Very rare
Mirabilite3
Na.$O,.l 0H20
Very common(?) Sepiolite
Mg,Si60,,(OH),.6H,0
Very rare
' Very common - mineral occurs in greater than 75% of the samples analyzed. Common -mineral occurs in 25-75%of thesample:
analyzed. Rare - mineral occurs in 10-25% of the samples analyzed. Very rare - mineraloccurs in less than 10% of the sample:
analyzed. These characterizations apply only to occurrence within the core and are not meant to imply any informationabout thc
relatlve abundance of the particular mineral. For example, some of the phases that are listed as being rare or very rare can
nonetheless, be relatively abundant in a pa~llcularsample, and minerals that may be listed as common can have very low
abundances.
Includes hydromagnesite and pseudohydromagnesite.
Because of Its extreme instablllty and very rapid dehydration, mirabilite was rarely identifiedby X-ray diffraction.

Y. Shang and W.M. Last

induration of the salts, due mainly to slight variations in mud


and organic matter content, there is an absence of visible bedding in the 23 m of recovered section.
The inorganic mineral suite in North Ingebrigt lake comprises three genetic types: allogenic (i.e. derived from weathering and erosion of the surrounding glacial drift, soils, and
bedrock in the watershed), endogenic (i.e, originating from
processes occurring within the water column of the lake), and
authigenic (i.e. originating from processes occurring once the
sediment has been deposited). The endogenic+authigenic
fraction dominates the sedimentary sequence (average 88%;
Fig. 3; Table 1) and consists of a complex assemblage of
mainly hydrated Na-, Mg-, (Na+Mg)-, and Ca-sulphate minerals, with smaller amounts of Mg-, (Ca+Mg)-, Na-, and
Ca-carbonate and -chloride minerals (Table 2). Although the
entire stratigraphic section is dominated by sulphate minerals, the relative proportion of Na-sulphates decreases upward
corresponding to a complementary increase in Mg- and
Ca-sulphates. Most of the endogenic and authigenic carbonate minerals show little stratigraphic variation. Magnesium
carbonates (magnesite, hydromagnesite, and pseudohydromagnesite) are most abundant between about 8 and 18 m
depth, whereas the carbonates in the upper metre of section
are dominated by disordered, nonstoichiometric dolomite
(Fig. 4).

In contrast, the minor amount of allogenic sediment (average 12%) is composed of a relatively simple mixture of clay
minerals, feldspar minerals, carbonate minerals, and quartz
(Fig. 5). This allogenic fraction closely reflects the composition of glacial deposits in the surrounding drainage basin (cf.
David, 1964). These detrital components are most abundant
in the upper metre of the section and at 5-7 m and 8.5-10 m
depths in the core.
Linear correlation coefficient analysis of the mineralogical and lithostratigraphic data shows, as expected, that variations in the allogenic components are significantly correlated
(at a=0.005confidence level) with changes in organic matter
content, and inversely correlated with variations in most of
the endogenic and authigenic components.Within the
endogenic and authigenic minerals, the changes in sulphate
and carbonate fractions show an inverse correlation, the variation in sulphates and chlorides exhibit a strong positive correlation, and fluctuations in the Na-sulphate salts are
inversely correlated with the Ca- and Mg-sulphates.

6
0.1
5

..I

1
8..

Figure 4. Stratigraphic variation in percentages of Mg-calcite, protodolomite,


aragonite, Mg-carbonate minerals, Nu-sulphate minerals, Mg-sulphate minerals,
and Ca-sulphate minerals in North Ingebrigt lake core. Heavy horizontal dotted lines
delineate mineral facies (see Table I). Gaps in plots represent sections in which there
is missing core.

GSC Bulletin 534

Weathering
indices

"'"1 1
16

Missing core

F.~..... F.

1 8 F . ~ ~ ~ * ~ .
I ,
,
I ,
,

00.00..

0 0 . 0 0 . .

1 ll
7-

. .F ~ . ~ * V * *FF.~.~...
~.*.

:i;~,i;~;;,

0.0.0..

0.0.0..

, ,

0 I02030

I02030
%

30

60

30
%

60

30
%

0.0

0 0 . 0 0 . .

24
60

0.0.0..

Intensity

Figure 5. Stratigraphic variation in percentages of quartz, K-feldspar, plagioclase,


total clay minerals, and allogenic carbonate minerals in North Ingebrigt lake core.
The carbonate weathering index is the ratio of detrital dolomite to calcite, and the
siliciclastic weathering index is the ratio of total detrital siliciclastic minerals to total
detrital carbonate minerals. An increase in these ratios suggests more intense
weathering within the watershed. Heavy horizontal dotted lines delineate mineral
facies (see Table 1). Gaps inplots represent sections in which there is missing core.

Mineral "facies
Based mainly on the OCCu~~-ence
and relative abundance of
endogenic and authigenic mineral phases in the core, the
recovered stratigraphic section from the lake was divided into
six mineralogical facies by means of average linkage cluster
analyses using a distance coefficient of greater than 0.2
(Joreskog et al., 1976; Beaumont and ~ a t r e l l ,1982; SAS
Institute Inc., 1989). It must be emphasized that the contacts
between these mineralogical facies are not visually distinctive and are commonly gradational over several tens of centimetres. Table I summarizes the average major mineral
composition of these six facies. The stratigraphic positions of
these facies are shown in Figures 3-5.
Although sulphate and carbonate minerals are the dominant soluble and sparingly soluble components in the inorganic fraction of North Ingebrigt lake sediment, the specific
mineralogy of these endogenic and authigenic precipitates is

complex (Table 2). Further subdivision of the mineral facies


into mineral zones was based on the detailed composition of
the samples in each facies. A total of 31 individual zones were
recognized in the 23
long core, with each zone distinguished by a specific suite of inorganic components
(Table 3).

Facies 1 - Na-sulphate (17.6 m to base of core,


3.44.5 m, and 0.8-1.6 m depth)
The most commonly occurring mineral facies in North
Ingebrigt lake is characterized by a relatively simple mineral
assemblage comprising mainly Na-sulphate salts with only
minor amounts of gypsum, magnesite, protodolomite,
aragonite, and halite. This facies is also identified by the presence of nitrate salts. The ratio of (endogenic+authigenic) to
allogenic components is very high, and the organic matter
content extremely low (Table 1).

Y . Shang and W.M. Last

Table 3. Stratigraphic sequence of mineral facies and mineral zones in North lngebrigt lake core.

Mineral
facie.

Mineral
zone

1 1
2

I1

Demh l m l
0.1-0.4
0.4-0.8
0.8-1.2

Salt mineral assemblages


Maior

1.2-1.6

Protodolomite. thenardite
Gypsum, thenardite

1 ~g-calcite- 1
/

asuphate

4
Muddy
Na-sulphate

Bischofite

Gypsum, Mg-calcite

Protodolomite, thenardite

1 Thenardite

Minor & trace


Ankerite, benstonite, despujolsite,
douglasite, leightonite, minrecordite

Gypsum, Mg-calcite

I Bloedite, Mg-calcite,

I Hanksite, nlter

2.4-3.4

3.4-3.9

1 Gypsum, thenardite
/ Thenardite

II

Gypsum, Mg-calcite

1I

Ca-sulphate:a-sulphate

/ Aragonite, gypsum,

( Hanksite, tychite

3.9-4.1

Thenardite

Gypsum, protodolomite

Niter

Thenardite

Protodolomite

Hanksite, niter

4.5-4.9

Thenardite

Mg-calcite, magnesite,
protodolomite

Hanksite

4.9-5.6

Gypsum, thenardite

Aragonite, Mg-calcite,
protodolomite

Arcanite, hanksite, inderborite

5.6-6.2

Thenardite

Gypsum, magnesite,
protodolomite

Arcanite, gaylussite, hanksite,


hexahydrite, sepiolite, wattevilleite

Gypsum, thenardite

Aragonite, protodoiomite

Arcanite, hanksite, polyhalite, sepiolite

1 1

6.2-7.2

Burkeite, epsomite, hanksite, leonhardtite,


mercallite, gorgeyite, sepiolite

8.2-9.4

Gypsum, thenardite

Aragonite, gypsum,
magnesite, protodolomite

9.4-10.2

Gypsum, magnesite,
thenardite

Aragonite, halite, Mg-calcite, Hanksite, mercallite, polyhalite, gorgeyite,


rapidcreekite, sepiolite
protodolomite

Gypsum, thenardite

Aragonite, magnesite,
protodolomite
Aragonite, magnesite,

10.5-1 1.2

11.2-1 1.7

Gypsum, thenardite

Aragonite, halite, magnesite, Hanksite, rapidcreeksite, sepiolite


Mg-calcite, protodoiomite

11.7-13.4

Gypsum, magnesite,
thenardite

Aragonite, protodolomite

Hanksite

Gypsum, magnesite,
thenardite

Aragonite, bloedite, halite,


Mg-calcite, protodolomite

Arcanite, despujolsite, eugsterite,


hanksite, rapidcreekite

10.2-10.5

Carnallite, hanksite, polyhalite

1 1

MgcarbonateNa-sulphate

Gypsum, thenardite

1 protodolomite

Hanksite, mercallite. polyhalite


Hanksite, leonhardtite, polyhaiite

14.1-15.1

Gypsum, protodolomite,
thenardite

Aragonite, magnesite

Hanksite, gorgeyite, leonhardtite

15.1-15.8

Gypsum, thenardite

Aragonite, magnesite,
protodolomite

Gayiussite

15.8-16.4

Magnesite, thenardite

Aragonite, protodolomite

Bonshtedtite, bradleyite, hanksite, pyrite

16.4-1 7.4

Gypsum, magnesite,
protodolomite, thenardite

Mg-calcite

Hanksite, leonhardtite, rapidcreekite

Hanksite

II

protodolomite
4.1-4.5

1
Na-sulphate

Arcanite, artinite, hanksite

Aragonite, bloedite, gypsum, Arcanite, despujosite, hanksite,


magnesite, protodolomite
leightonite. mercaite, rapidcreekite
Mg-calcite.
Benstonite, epsomite, gbrgeyite,
hanksite,
. mercallite, polyhalite

protodoiom~e

Na-sulphate

Ancillary

Bloedite, protodolomite,
thenardite

1
5

17.6-19.5

Thenardite

Magnesite, protodoiomite

19.5-20.6

Thenardite

Halite

Hanksite, nitratine, nitrobarite

20.6-21.I

Thenardite

Gypsum

Nitratine, protodolomite

21.l-22.2

Thenardite

Gypsum

, 22.2-23.0

Thenardite

Gypsum

Facies 2 - Mg-carbonate-Na-sulphate
(11.7-17.6 m depth)
Facies 2 is recognized by a relatively high content of
magnesite, aragonite, and gypsum (Table 1). Both magnesite
and gypsum increase upward over the 6 m of stratigraphic
section in which this facies occurs. Facies 2 is also distinguished by the presence of pyrite and a wide array of ancillary
sulphate minerals and carbonate-sulphate double salts.

Hanksite, nitratine

Facies 3 - Ca-sulphate-Na-sulphate (8.2-11.7 m)


Facies 3 has the highest gypsum content in the stratigraphic
sequence and is also characterized by very low amounts of
Mg-sulphate salts (Table 1). Both the organic matter and the
allogenic fraction of the sediment show an upward increase in
abundance over the 2.5 m of core in which this facies is found.

GSC Bulletin 534


Facies 4 - muddy Na-sulphate (4.5-8.2 m depth)
Facies 4 is the most detrital-rich facies of the sedimentary
sequence, with nonevaporite minerals constituting up to 50%
of the sediment. The endogenic and authigenic precipitates
are mainly Na- and Ca-sulphates, with relatively low proportions of carbonates and no Mg-sulphate salts (Table 1).
Facies 5 - muddy Mg-calcite-Na-sulphate
(1.6-3.4 m depth)
Facies 5 also has a relatively high content of detrital clastic
minerals and associated medium to dark grey colour. The
endogenic+authigenic fraction is further characterized by a
high proportion of Mg-calcite, gypsum, and magnesite, and
an absence of Mg-sulphates (Table 1).The salts of facies 5
are less well indurated than those in the other parts of the
section.
Facies 6 - dolomitic Na-sulphate (0.1-0.8 m depth)
The least commonly encountered facies in the stratigraphic
record is marked by relatively high proportions of disordered,
nonstoichiometiic dolomite associated with a complex mixture of Na-, (Na+Mg)-, and Ca-sulphate salts and halite
(Table 1). Facies 6 is only found in the upper metre of the section. Like facies 5, it is poorly indurated and dark coloured.

Chronostratigraphy
The chronostratigraphy of the North Ingebrigt lake salt
sequence is poorly constrained owing to a paucity of
material suitable for dating. Large quantities of core were
processed and examined in an unsuccessful attempt to find
datable, well preserved upland and shoreline macrofossils.
Therefore, finely disseminated organic matter from two
samples was used to establish a preliminary chronostratigraphy for the recovered sedimentary sequence. A 14C
age of 5544 f 6 6 BP (5912 calendar years before present)
was determined on a sample from the base of facies 5 at
3.2-3.4 m depth, and a sample from near the base of the core
(22.1-22.9 m) yielded an age of 10 250 f 150BP (12 181 calendar years before present). In addition, a sample of
endogenic carbonate material from facies 3 at the base of a
9 m long core in the southern part of the basin yielded a date
of 8240 +I20 BP (9176 calendar years before present). Further temporal control is provided by the presence of a borate
mineral zone (inderborite) at 5.6 m depth, which may be an
indication of a diagenetically altered volcanic ash corresponding to the 6800 BP Mazama eruption. Petrographic
and geochemical analysis of this borate zone is continuing,
but it must be emphasized that the interpretation of this zone
as an altered volcanic ash is still equivocal.

DISCUSSION
General paleolimnology of North Ingebrigt lake
The 23 m of salt-rich sediment recovered from the North
Ingebrigt lake basin was deposited in a saline to hypersaline
environment. The predominance of Na-sulphate salts and
paucity of nonevaporitic sediments indicate that a saturated
~ in- the basin for
brine, high in Na+ and ~ 0ions,~existed
nearly all of the Holocene. However, the complex assemblage of evaporites, including a diverse array of Ca-, Mg-,
Na-, and K-sulphates and -carbonates, suggests that the lake
experienced considerable short-term fluctuation in brine
composition. The common occurrence of chlorides and Mgand K-sulphates in the evaporitic deposits of the lake implies
that the salinity of the water periodically reached, and probably exceeded, values of 350 g . ~ - l T D sIn
. contrast, the presence of Mg-calcite and aragonite indicates that lower salinity,
Ca-bicarbonate waters were also present. Dramatic fluctuations in brine chemistry and salinity such as these are not
unusual in modern saline playas of the region. For example,
Ceylon Lake in south-central Saskatchewan has, over the past
several decades, ranged from a freshwater Ca-bicarbonate
lake (TDS of less than 5 g . ~ - lto) a sulphate and chloride mineral precipitating basin with a salinity of nearly 400 g . ~ - l
TDS (Last, 1987,1989b, 1990).
Although many saline lakes and playas in the northern
Great Plains of western Canada and elsewhere are characterized by relatively high levels of organic productivity
(Warren, 1986; Hammer, 1986; Slezak and Last, 1987;Lyons
et al., 1994), the sediments recovered from North Ingebrigt
lake are remarkably low in organic content. This scarcity of
organic matter in the stratigraphic record is most likely a
reflection of a combination of factors, including 1) high
salinities experienced in the lake, which curtail both organism diversity and productivity; 2) strongly oxidizing conditions at the sediment-water interface; and 3) only small
amounts of terrestrial organic debris reaching the centre of
the basin due to negligible stream flow.
Changes in the depth of brine and long-term fluctuations
in water level in the North Ingebrigt basin undoubtedly
occurred over the past 10 000 years, but are difficult to assess
from the preserved lithostratigraphic record. In contrast to the
salt-dominated stratigraphic records in nearby Ingebrigt,
Chappice, and Freefight lakes, and Little Manitou Lake in
central Saskatchewan (Ruffel, 1968; Last, 1993b; Vance
et al., 1993; Sack and Last, 1994), there is no compelling
petrographic, lithostratigraphic, or geochemical evidence
that North Ingebrigt lake was ever a deep-water basin.
Indeed, shallow-water hypersaline playa conditions, probably similar to those of today, existed throughout most of the
Holocene. However, the absence of pedogenic horizons, recognizable erosional and solution-modified contacts and bedding planes, and relatively coarse-grained allogenic sediment
suggests that the basin probably never experienced significant or extended periods of complete dryness and subaerial
desiccation.

Y. Shang and W.M. Last

Although absolute age dating of the North Ingebrigt lake


record has provided only a very preliminary chronostratigraphic framework, it nonetheless suggests that there
have been considerable changes in sediment accumulation
rates during the Holocene. Deposition of the sediments of
facies 1 and facies 2 in the lower half of the sequence appears
to have been rapid, with accumulation rates approaching
50 cm.l00a-~.During the mid- and late Holocene, this rapid
rate of salt accumulation decreased significantly, and extrapolated linear sedimentation rates averaged an order of magnitude less than those exhibited in the early part of the record.

Geochemical conditions of sedimentation


A basic tenet of chemical sedimentology is that, in a
hydrologically closed system undergoing evaporative concentration, a complete assemblage of nondetrital minerals
must have the same bulk ionic composition as the original
water (Hardie, 1984). As discussed elsewhere (Wasson et al.,
1984; Torgersen et al., 1986; Teller and Last, 1990; Last,
1994), there are many assumptions and limitations that must
be accepted in the application of this straightforward principle to paleolimnology. Nonetheless, North Ingebrigt lake, a
topographically and hydrologically closed basin having a relatively small catchment and a long and apparently continuous
record of chemical sedimentation from hypersaline brines, is
ideally suited for paleochemical reconstructions based on
these types of thermodynamic and mass balance calculations.
Our ability to deduce the composition of a brine from the
preserved mineral record in a lacustrine basin has been
improved by significant advances in thermodynamic models
of salt dissolution and mineral precipitation from concentrated solutions (Harvie and Weare, 1980;Harvie et al., 1982,
1984; Greenberg and Moller, 1989; Spencer, et al., 1990).
The theoretical framework for this brine reconstruction follows discussions by Garrels and Christ (1965), Eugster and
Smith (1965), Wood (1975), Smith (1979), and Wasson et al.
(1984). Figure 6 shows changes in the relative chemical
activities of the major dissolved components in the precipitating brine in North Ingebrigt lake through the Holocene.
The salts of facies 1 (Na-sulphate facies) were precipitated from a geochemically simple brine with a cation content
dominated by sodium and the anions comprising mainly sulphate and nitrate. The source for the increased relative chemical activity of nitrate (the chemical activity of a major
dissolved component is indicated by a superscript 'a' preceding the formula for the component, thus aN03-)in the precipitating brine compared to the rest of the Holocene stratigraphic
sequence is not known with certainty. Rutherford (1967)
reported elevated NOg-levels in some of the shallow (glacial
drift) groundwater of the area. Shang (1999) showed that
evaporative concentration of these glacial drift groundwaters
can result in niter and soda niter precipitation. Similarly, it is
well known that oxidation of organic matter in a strongly
alkaline environment under relatively arid conditions can
result in the formation of Na- and K-nitrate salts (Clarke,
1924).

During deposition of facies 2 (Mg-carbonate-Nasulphate facies) between about 9200 and 8000 BP, the relative chemical activities of ca2+, Na+, and HC03- increased,
aS042-remained high, and aN03 - gradually decreased. The
resulting brine was thus considerably more complex than that
which occupied the basin during the first 1000-1500 years of
the record. This facies marks the earliest significant occurrence of carbonate in the lake water. The presence of
gaylussite rather than pirssonite suggests relatively low temperatures, high pH, and high aH20 conditions. The occurrence of eugsterite, a (Na+Ca)-sulphate, together with
gypsum and bloedite, indicates a precipitating solution with a
varying but generally very high molar Na:Ca ratio
(Vergouwen, 1981; Van Doesburg et al., 1982).
The salts of facies 3 (Ca-sulphate-Na-sulphate facies)
indicate deposition in a brine with higher a ~ a 2values
+
than
existed previously. The widespread occurrence of minor
amounts of hanksite and polyhalite, interpreted as
penecontemporaneous diagenetic products of gypsum, suggests that this brine was commonly elevated in C1- and K+,
and was highly alkaline.
Deposition of facies 4 (muddy Na-sulphate facies)
between about 7300 and 6200 BP and facies 5 (muddy
Mg-calcite-Na-sulphate facies) between 5600 and 2900 BP
represents significant freshening episodes of the brine. However, the lake still remained saline to hypersaline. The a ~ a 2 +
and aHC03- values show an increase relative to facies 1-3,
and there are complementary decreases in a ~ g 2 +aNa+,
,
and
aC1- coincident with brine dilution.
One of the most dramatic changes in water composition in
the North Ingebrigt record occurred about 1000 years ago.
Although the transition from facies 5 to facies 4 was gradual
over the period ca. 4000-2000 BP, the chemical conditions
responsible for facies 4 ended abruptly ca. 1000 years ago.
After this time, there was a major increase in relative a ~ g 2 +
and aHC03-values and a decrease in a~042-,
a ~ a 2 +and
, aNa+.
Although this uppermost dolomitic Na-sulphate facies may
represent an overall freshening of the brine, hypersaline conditions remained common, as indicated by the occurrence of
halides and Mg-sulphates. Facies 6 is unusual, with no known
modem analogues among today's saline playas of western
Canada.
Any discussion of water-level fluctuations or lake-depth
changes in a basin that annually or periodically undergoes
complete drying is meaningless. Likewise, it is difficult to
evaluate or summarize brine solute concentration changes in
a water body that regularly experiences great ranges of salinity associated with playa flooding and desiccation. However,
the mineralogical record does provide some important information about the changes in mean atmospheric relative
humidity. As discussed elsewhere (Harbeck, 1955;
Langbein, 1961; Hosler, 1979), evaporation and evaporative
concentration of a water body leading to mineral precipitation
is a complex process. The effective, long-term average atmospheric relative humidity of the location plays a pivotal role in
helping to determine the ultimate suite of evaporites that is

GSC Bulletin 534

Mineral
facies
6
1

11000

Low

High Low

HCO,

Low

High Low

High

Low

High

Relative chemical activity


SO,

HighLow

CI

HighLow

Mineral
facies

NO,

HighLow

High

Relative chemical activity


Figure 6. Relative chemical activities of major dissolved (A) cation
and ( B )anion components of lake (or interstitial) brines versus time,
based on the endogenic+authigenic mineral suite in North Ingebrigt
lake. Relative activity values increase to the right; heavy horizontal
dashed lines delineate mineral facies (see Table I). Temporal scale is
in 14cyears BP and was established by simple linear interpolation.

Y. Shang and W.M.

formed (Kinsman, 1976; Sonnenfeld, 1984). Regardless of


the actual mechanism of evaporite mineral formation, the
presence of a particular mineral or suite of minerals is an indication that the brine was thermodynamically saturated with
respect to that inorganic precipitate. The chemical activity of
the water (aH20)in the precipitating solution is directly proportional to the water vapour pressure of the atmosphere,
which in turn is related to the mean atmospheric relative
humidity (Kinsman, 1976). Thus, the wealth of experimental
and theoretical data on thermodynamic conditions of salt
mineral formation (Eugster and Smith, 1965; Bradley and

Relative hurr~idity

+$$a5

\&

0 ,

Last

Eugster, 1969; Hardie and Eugster, 1970; Braitsch, 1971;


Smith and Stuiver, 1979; Eugster et al., 1980) can be used to
estimate aH20,and hence the relative humidity of the basin.
During the periods 10 350-8700 BP, 6500-5500 BP, and
1500-1000 BP, the North Ingebrigt lake basin experienced
climatic conditions in which the mean relative humidity was
considerably lower than that of today (Fig. 7). In contrast,
comparatively humid conditions (but not wet enough to cause
a basic change in the hydrological setting of the playa) are
suggested by higher relative humidity during the periods
7200-6500 BP and 5500-3000 BP. Finally, the millennium
between 8000 and 7000 BP was characterized by considerable variability, with rapid, short-term fluctuations between
high and low relative humidity.

North Zngebrigt lake relative to other Holocene


lacustrine records
Although it is not our intent in this paper to reconstruct the
regional climatic history, it is nonetheless tempting to compare the changes in mean atmospheric relative humidity interpreted for North Ingebrigt lake with paleolimnogical
fluctuations at other sites in the region with long Holocene
records (Table 4).

11 000

Low

High

Figure 7. Variation in mean atmospheric relative humidity at


North Ingebrigt lake as estimated by the evaporitic mineral
suite. Numbers on the right refer to the mineral facies (see
Table I ) . Temporal scale is in 14cyears BP and was
established by simple linear interpolation.

The 7500-year long sediment record from Chappice Lake,


Alberta, a hypersaline playa located about 80 km west of
North Ingebrigt lake, is one of the longest and best documented lacustrine paleohydrological records in the entire
northern Great Plains region (Vance et al., 1992,1993,1995).
Although temporal control of the North Ingebrigt lake record
is much less tightly constrained than that of Chappice Lake,
broad similarities exist. Both basins show an early Holocene
episode of significant fluctuation in relative humidity and in
water level and salinity conditions (prior to 6000 BP at
ChappiceLake; 8500-7000 BP at North Ingebrigt lake). Similar early Holocene periods of rapidly changing conditions,
from high to low water levels and from hypersaline to fresh
water, were noted in stratigraphic records from Ceylon lake,
about 400 km southeast of Ingebrigt Lake (Last, 1990), and
from Clearwater Lake, 100 km to the northeast (Last et al.,
1998). Similarly, a mid-Holocene interval of low-water,
saline to hypersaline conditions correlative to the very low
mean atmospheric relative humidity at 6000-5000 BP at
North Ingebrigt lake has also been documented at Chappice
and Ceylon lakes as well as several other lacustrine sites in the
Prairie region (Vance et al., 1995). Finally, the lengthy period
of stable and relatively moist conditions between about 5000
and 2000 BP is recorded in numerous other basins on the
northern Great Plains (see overviews in Lemmen, 1996; Xia
et al., 1997; Vance, 1997).

GSC Bulletin 534

Table 4. Comparison of mean relative humidity at North lngebrigt lake to interpreted water levels
and salinities at Chappice Lake, Alberta, and Clearwater and Ceylon lakes, Saskatchewan.
"C age
(years BP)

Clearwater Lake Water level and


salinity

North lngebrigt lake Mean relative humidity

Chappice Lake Water level and


salinity

High

Fluctuation from low to high


and fresh to saline

Present
1000

Low and hypersaline


High and fresh water

Low water
Stable conditions deep and fresh water

Lnw
2000
Deep and relatively fresh water

Very low
,nlyrl

6000
7000

t:-L

Repeated fluctuation from


low to high

Missing section

High
9000
10000

Stable conditions low water and


hypersaline

Repeated fluctuation from dr!


to deep and fresh to
hypersaline
Base of record

Very low to low

Stable conditions low water and


hypersaline
High and fresh water -

Stable conditions low water and hypersaline

5000

Ceylon lake Water level and


salinity

Deep and fresh water


Deep to shallow (dry at 8200 BP)
and saline to hypersaline
Shallow and hypersaline
Shallow and fresh water saline

Very low to dry


and hypersaline

Repeated fluctuation
from shall0w deep
and fresh to
hypersaline

Deep and fresh water

CONCLUSION

REFERENCES

There are many salt-dominated playa basins in the Great


Plains of North America whose sediment records are conducive to paleolimnological study. North Ingebrigt lake is one
of the first of these lakes to undergo detailed mineralogical
and lithostratigraphic investigations. Obviously, deciphering
all the intricate processes that control the relationships
between lake hydrology, regional climate, and the chemistry
of a brine is an arduous task even in a small, presumably
rather simple evaporating pan like North Ingebrigt lake. It has
been repeatedly shown that there is relatively little relationship between brine chemistry of modern lakes in the northern
Great Plains and their specific climatic setting (e.g. Last,
1989c, 1992; Lambert, 1989; LaBaugh and Swanson, 1992).
This suggests that any linkage between water composition
and climate on a regional basis is extremely tenuous. Indeed,
the Holocene mineralogical record preserved in North
Ingebrigt basin emphasizes the complexity, rather than the
simplicity, of long-term brine composition changes and the
interplay between the many intrinsic processes that are operating within the basin itself and the various extrinsic factors
that help control the geochemical aspects of the lake.

Assarsson, G.O.
1950a: Equilibria in aqueous systems containing K, Na, Ca, Mg and CI,
Part I. The ternary system CaC12-KC1-H20; Journal of American
Chemistry Society, v. 72, p. 1433-1436.
1950b: Equilibria in aqueous systems containing K, Na, Ca, Mg and Cl,
Part LI. The quaternary system CaC12-KC1-NaC1-H20; Journal of
American Chemistry Society, v. 72, p. 1437-1441.
1950c: Equilibria in aqueous systems containing K, Na, Ca, Mg and C1,
Part 111. The ternary system CaC12-MgC12-H20; Journal of
American Chemistq Society, v. 72, p. 1442-1444.
Beaumont, J.R. and Gatrell, A.C.
1982: An introduction to Q-analysis; Geo Abstracts, Norwich, United
Kingdom, Concepts and Techniques in Modern Geography, No. 34,
55 p.
Berglund, B.E. (ed.)
1986: Handbook of Holocene Palaeoecology and Palaeohydrology; John
Wiley and Sons, New York, 869 p.
Bradley, W.H. and Eugster, H.P.
1969: Geochemistry and paleolimnology of the trona deposits and associated authigenic minerals of the Green River Formation of Wyoming; United States Geological Survey, Professional Paper 496-B,
p. 71.
Braitsch 0.
1971: Salt Deposits, Their Origin and Composition; Springer-Verlag,
New York, 297 p.
Canadian National Committee for the
International Hydrologic Decade
1978: Hydrologic Atlas of Canada; Fisheries and Environment Canada,
Ottawa, 75 p.
Christiansen, E.A.
1979: The Wisconsinan deglaciation of southern Saskatchewan and adjacent areas; Canadian Journal of Earth Sciences, v. 16, p. 913-938.
Clarke, F.W.
1924: The data of geochemistry; United States Geological Survey
Bulletin, v. 770, p. 871.
Cole, L.H.
1926: Sodium sulphate of western Canada - occurrence, uses and technology; Canada Department Mines, Publication 646, 155 p.
David, P.P.
1964: Surficial geology and ground water resources of the Prelate area
(72K); Ph.D. thesis, McGill University, MontrBal, QuBbec, 329 p.
Davis. R.B. (ed.)
. ,
1990: Paleolimnology and the Reconstruction of Ancient Environments;
Kluwer Academic Publishers, Dordrecht, The Netherlands, 254 p.

ACKNOWLEDGMENTS
This research was made possible through funding from the
Geological Survey of Canada's Palliser Triangle Global
Change Project, the Natural Sciences and Engineering
Research Council of Canada (NSERC), and the University of
Manitoba. Anne Ross and Grant Caresle provided valuable
assistance in the laboratory. We thank J. Donovan,
R.A. Klassen, and D.S. Lemmen for their helpful comments
and suggestions on an earlier draft of this manuscript.

Y. Shang and W.M. Last


Dean, W.E.
1974: Determination of carbonate and organic matter in calcareous sediments and sedimentary rocks by LOI; Journal of Sedimentary
Petrology, v. 44, p. 242-248.
Dearing, J.A.
1997: Sedimentary indicators of lake-level changes in the humid temperate zone: a critical review; Journal of Paleolimnology, v. 18,
p. 1-14.
Epp, H.T. and Townley-Smith, L.
1980: The Great Sand Hills of Saskatchewan; report prepared for
Saskatchewan Department of the Environment, Regina, 156 p.
Eugster, H.P.
1971: The beginning of experimental petrology; Science, v. 173,
p. 481-489.
Eugster, H.P. and Smith, G.I.
1965: Mineral equilibria in the Searles Lake evaporites, California;
Journal of Petrology, v. 6, p. 473-522.
Eugster, H.P., Hawie, C.E., and Weare, J.H.
1980: Mineral equilibria in the six-component seawater system,
Na-K-Mg-Ca-S04-CI-H20,at 250C; Geochimica et Cosmochimica
Acta, v. 44, p. 1335-1347.
Evans, M.S.
1993: Paleolimnological studies of saline lakes; Journal of
Paleolimnology, v. 8, p. 97-102.
Garrels, R.M. and Christ, C.L.
1965: Solutions, Minerals and Equilibria; Harper and Row, New York,
450 p.
Gell, P.A., Barker, P.A., De Deckker, P., Last, W.M., and Jelicic, L.
1994: The Holocene history of West Basin Lake, Victoria, Australia:
chemical changes based on fossil biota and sediment mineralogy;
Journal of Paleolimnology, v. 12, p. 235-258.
Gierlowski-Kordesch, E. and Kelts, K.
1994: Introduction; in Global Geological Record of Lake Basins, Volume
1, (ed.) E. Gierlowski-Kordesch and K. Kelts; Cambridge
University Press, New York, p. xvii-xxxiii.
Goldsmith, J.R. and Graf, D.L.
1958: Relation between lattice constants and composition of the Ca-Mg
carbonates; American Mineralogist, v. 43, p. 84-101.
Greenberg, J.P. and Mrrller, N.
1989: The predicticm of mineral solubilities in natural waters: a chemical
equilibrium model for the Na-K-Ca-CI-S04-Hz0 system to high
concentrations for 0-2500C; Geochimica et Cosmochimica Acta,
V.53, p. 2503-2518.
Hammer, U.T.
1986: Saline Lake Ecosystems of the World; Dr. W. Junk Publishing,
Dordrecht, The Netberlands, 616 p.
Harbeek, G.E.
1955: The effect of salinity on evaporation; United States Geological
Survey, Professional Paper 272-A, p. 470477.
Hardie, L.A.
1968: The origin of the Recent non-marine evaporite deposit of Saline
Valley, Inyo County, California; Geochimica et Cosmochimica
Acta, v. 41, p. 833-834.
1984: Evaporites: marine or non-marine?; American Journal of Science,
V.284, p. 193-240.
Hardie, L.A. and Eugster, H.P.
1970: The evolution of closed-basin brines; Mineralogical Society of
America, Special Paper, v. 3, p. 273-290.
Hardy, R. and Tucker, M.
1988: X-ray powder diffraction of sediments; in Techniques in
Sedimentology, (ed.) M. Tucker; Blackwell Scientific Publishing,
Boston, Massachusetts, p. 191-228.
Harvie, C.E. and Weare, J.H.
1980: The prediction of mineral solubilities in natural waters: the
Na-K-Mg-Ca-CI-S04-Hz0 system from zero to high concentrations at 25C; Geochimica et Cosmochimica Acta, v. 44,
p. 981-997.
Harvie, C.E., Eugster, H.P., and Weare, J.H.
1982: Mineral equilibria in the six-component seawater system,
Na-K-Mg-Ca-CI-S04-H20 at 250C. II: compositions of the saturated solutions; Geochimica et Cosmochimica Acta, v. 46,
p. 1603-1618.
Harvie, C.E., Mbller, N., and Weare, J.H.
1984: The prediction of mineral solubilities in natural waters: the Na-K-MgCa-H-CIS04-OH-HCO~-CO3-C02-H~0
system to high ionic strengths
at 250C; Gwchimica et Cosmochimica Acta, v. 48, p. 723-751.

Holser, W.T.
1979: Mineralogy of evaporites; in Marine Minerals, (ed.) R.G. Burns;
Mineralogical Society of America, Short Course Notes, v. 6,
p. 21 1-294.
Joreskog, K.G., Klovan, J.E., and Reyment, R.A.
1976: Geological Factor Analysis; Elsevier Scientific Publishing
Company, New York, 178 p.
Kinsman, D.J.J.
1976: Evaporites: relative humidity control of primary facies; Journal of
Sedimentary Petrology, v. 46, p. 273-279.
LaBaugh, J.W. and Swanson, G.A.
1992: Changes in chemical characteristicsof water in selected wetlands in
the Cottonwood Lake area, North Dakota, U. S. A,, 1967-1989; in
Aquatic Ecosystems in Semi-arid Regions: Implications for
Resource Management, (ed.) R.D. Robarts and M.L. Bothwell;
Environment Canada, National Hydrology Research Institute,
Symposium No. 7, p. 149-162.
Lambert, S.
1989: Hydrogeochemistry of saline lakes in the northern Interior Plains of
westenl Canada and northern United State; B.Sc. thesis, University
of Manitoba, Winnipeg, Manitoba, 162 p.
Langbein, W.B.
1961: Salinity and hydrology of closed lakes; United States Geological
Survey, Professional Paper 412, p. 1-20.
Last, W.M.
1980: Sedimentology and postglacial history of Lake Manitoba; Ph.D.
thesis, University of Manitoba, Winnipeg, Manitoba, 686 p.
1984: Sedimentology of playa lakes of the northern Great Plains;
Canadian Journal of Earth Sciences, v. 21, p. 107-125.
1987: Sedimentology,geochemistry, and evolution of a saline playa in the
northern Great Plains of Canada; in Geochemistry and Mineral
Formation in the Earth Surface, (ed.) R. Rodriguez-Clemente and
Y. Tardy; Consejo Superior de Investigaciones Cientificas, Centre
National delaRecherche Scientifique,Madrid, Spain, p. 319-337.
1989a: Continental brines and evaporites of the northern Great Plains of
Canada; Sedimentq Geology, v. 64, p. 207-221.
1989b: Sedimentology of a saline playa in the northern Great Plains,
Canada; Sedimentology, v. 36, p. 109-123.
1989c: Salt lakes of western Canada: a spatial and temporal geochemical
perspective; in Proceedings of the Symposium on Water
Management Affecting the Wet-to-Dry Transition: Planning at the
Margins, (ed.) W. Nicholaichuk and H. Steppuhn; Water Studies
Institute, University of Regina, Regina, Saskatchewan, p. 99-1 13.
1990: Paleochemistry and paleohydrology of Ceylon lake, a
salt-dominated playa basin in the northern Great Plains, Canada;
Journal of Paleolimnology, v. 4, p. 219-238.
1993a: Salt dissolution features in salinelakes of the northern Great Plains,
western Canada; Geomorphology, v. 8, p. 321-334.
1993b: Geolimnology of Freefight Lake: an unusual hypersaline lake in the
northern Great Plains of western Canada; Sedimentology, v. 40,
p. 431-448.
1994: Paleohydrology of saline playas in western Canada: perspectives
from Palliser's Triangle; in Paleoclimate and Basin Evolution of
Playa Systems, (ed.) M. Rosen; Geological Society of America,
Special Paper 289, p. 69-8 I.
1996: Bulk composition, texture, and mineralogy of Lake Winnipeg core
and surface grab samples; inLake Winnipeg Project: Cruise Report
and Scientific Results, (ed.) B.J. Todd, C.F.M. Lewis,
L.H. Thorleifson, and E. Nielsen; Geological Survey of Canada,
Open File 3 113, p. 209-220.
Last, W. M. and Slezak, L.A.
1987: Sodium sulphate deposits of western Canada: geology, mineralogy,
and origin; in Economic Minerals of Saskatchawan, (ed.)
C.F. Gilboy and L.W. Vigrass; Saskatchewan Geological Society,
Special Publication No. 8, p. 197-205.
Last, W.M., Vance, R.E., Wilson, S., and Smol, J.P.
1998: A multi-proxy record of early Holocene hydrologic change on the
northern Great Plains of southwestern Saskatchewan, Canada; The
Holocene, v. 8, p. 503-520.
Lemnien, D.S. (ed.)
1996: Landscapes of the Palliser Triangle, Guidebook for the Canadian
Geomorphology Research Group Field Trip; Canadian Association
of Geographers, Annual Meeting, Saskatoon, 92 p.
,

GSC Bulletin 534


Lyons, W.B., Hines, M.E., and Last, W.M.
1994: Sulfate reduction rates in sediments of salt lakes of differing chemistries: implications for organic carbon preservation; in
Sedimentary Records of Modern and Ancient Saline Lakes, (ed.)
R. Renaut and W.M. Last; Society of EconomicPaleontologistsand
Mineralogists,Special Publication No. 50, p. 13-21.
Marquart, R.G.
1986: pPDSM: mainframe searchlmatch on an IBM PC; Powder
Diffraction, v. 1, p. 34-36.
Multhauf, R.P.
1978: Neptune's Gift: a History of Common Salt; The Johns Hopkins
University Press, Baltimore and London, 325 p.
Plummer, L.N., Parkhurst, D.L., and Kosiur, D.R.
1975: MIX2: a computer program for modeling chemical reactions in natural waters; United States Geological Survey, Water-Resources
Inventory Report 75-76,73 p.
Plummer, L.N., Parkhurst, D.L., Fleming, G.W., and Dunkle, S.A.
1988: A computer program incorporating Pitzer's equations for calculation of geochemical reactions in brines; United States Geological
Survey, Water-Resources Invento~yReport 88-4153,310 p.
Rollins, L.
1989: PCWATEQ: a PC version of the water chemistry analysis program
WATEQF; Phoenix Technologies Ltd., San Jose, California.
Ruffel, P.G.
1968: Development of the largest sodium sulfate deposit in Canada;
Canadian Institute of Mining and Metallurgy Bulletin, v. 61,
p. 483-488.
Rutherford, A.A.
1967: Water quality survey of Saskatchewan groundwaters;
Saskatchewan Research Council, Report C-66-1,267 p.
1970: Water quality survey of Saskatchewan surface waters;
Saskatchewan Research Council, Report C-66-1,266 p.
Sack, L.A. and Last, W.M.
1994: Lithostratigraphyand recent sedimentation history of Little Manitou Lake, Saskatchewan,Canada;Journal of Paleolimnology,v. 10,
p. 199-212.
SAS Institute Inc.
1989: SASISTAT user's guide, release 6.03 edition; Cary, North
Carolina, 1028 p.
Saskatchewan Research Council
1985: Surficial geology of the Prelate area (72-K) Saskatchewan;
Saskatchewan Research Council, Preliminary Map, scale
1:250 000.
Schulb, L.G.
1964: Quantitative interpretation of mineralogical composition from
X-ray and chemical data for the Pierre Shale; United States
Geological Survey, Professional Paper 391-C, 31 p.
Shang, Y.
1999: Paleolimnology, paleochemistry and paleohydrology of North
Ingebright lake, southwestern Saskatchewan,Canada; Ph.D. thesis,
University of Manitoba, Winnipeg, Manitoba, 425 p.
Sheskin, D.J.
1997: Handbook of Parametric and Nonparametric Statistical Procedures;
CRC Press, New York, 718 p.
Slezak, L.A. and Last, W.M.
1987: Organic matter production and preservation in Freefight Lake,
Saskatchewan; Geological Association of Canada-Mineralogical
Association of Canada Annual Meeting, Program with Abstracts,
v. 12, p. 90.
Smith, G.I.
1979: Subsurface stratigraphy and geochemistry of Late Quaternary
evaporites, Searles Lake, California; United States Geological
Survey, Professional Paper 1043, 130 p.
Smith, G.I., and Friedman, I.
1986: Seasonal diagenetic changes in salts of Owens Lake, California,
1970-77; in Studies in Diagenesis, (ed.) F.A. Mumpton; United
States Geological Survey, Bulletin 1578, p. 21-29.
Smith, G.I. and Stuiver, M.
1979: Subsurface stratigraphy and geochemistry of Late Quaternary
evaporites, Searles Lake, California; United States Geological
Survey, Professional Paper 1043, p. 1-130.
Smoot, J.P. and Lowenstein, T.K.
1991: Depositional environments of non-marine evaporites; in
Evaporites, Petroleum and Mineral Resources, (ed.) J.L. Melvin;
Elsevier Scientific Publishing, New York, p. 189-348.

Sonnenfeld, P.
1984: Brines and Evaporites; Academic Press, New York, 613 p.
Spencer, R.J., Mgller, N., and Weare, J.H.
1990: The prediction of mineral solubilities in natural waters: a chemical
equilibrium model for the Na-K-Ca-CI-S04-HzO system to high
concentrations at temperatures less than 250C; Geochimica et
Cosmochimica Acta, v. 54, p. 575-590.
Stewart, F.H.
1963: Marine evaporites; United States Geological Survey, Professional
Paper 440-Y, 53 p.
Sullivan, T.J. and Charles, D.F.
1994: The feasibility and utility of a paleolimnology/paleoclimate data
cooperative for North America: Journal of Paleolimnologv,
-. v. 10,
p. 265-273.
Teller.. J.T.
and Last. W.M.
ical
in playas and salt lakes, with examples
1990: ~ a l e o h ~ d r o l o ~indicators
from Canada, Australia, and Africa; Palaeogeography,
Palaeoclimatology,Palaeoecology, v. 76, p. 215-240.
Tomkins, R.V.
1953: Magnesium in Saskatchewan; Saskatchewan Department of
Mineral Resources, Report 11, 23 p.
1954: Natural sodium sulphate in Saskatchewan; Saskatchewan
Department of Mineral Resources, Report 6.71 p.
Torgersen, T., DeDeckker, P., Chivas, A.R., and Bowler, J.
1986: Salt lakes: a discussion of processes influencing palaeoenvironmental intermetation and recommendations for future
work; Palaeogeography, Palaeoclimatology,Palaeoecology, v. 54,
p. 7-19.
Truesdell, A.H. and Jones, B.F.
1974: WATEQ, a computer program for calculating chemical equilibria
of natural waters; Journal of Research of the United States
Geological Survey, v. 2, p. 233-274.
Vance, R.E.
1997: The Geological Survey of Canada's Palliser Triangle Global
Change Project: a multidisciplinary geolimnological approach to
predicting potential global change impacts on the northern Great
Plains; Journal of Paleolimnology, v. 17, p. 3-8.
Vance, R.E. and Last, W.M.
1994: Paleolimnology and global change on the southern Canadian
Prairies; in Current Research 1994-B; Geological Survey of
Canada, p. 49-58.
Vance, R.E., Beaudoin, A.B., and Luckman, B.H.
1995: The paleoecological record of 6 ka BP climate in the Canadian
Prairie Provinces; Geographie physique et Quaternaire, v. 49,
p. 81-98.
Vance, R.E., Clague, J.J., and Mathewes, R.W.
1993: Holocene paleohydrology of a hypersaline lake in southeastern
Alberta; Journal of Paleolimnology, v. 8, p. 103-120.
Vance, R.E., Mathewes, R.W., and Clague, J.J.
1992: 7000 year record of lake-lcvel change on the northern Great Plains:
a high-resolution proxy ofpast climate; Geology,v. 20, p. 879-882.
Van Doesburg, J.D., Vergouwen, L., and Plas, L.V.D.
1982: Konyaite, NazMg(SOs)z5HzO, a new mineral from the Great
KonyaBasin,Turkey; Amcrican Mineralogist,v. 67,p. 1035-1038.
Vergouwen, L.
1981: Eugsterite, a new salt mineral; American Mineralogist, v. 66,
p. 632436.
Warren, J.K.
1986: Source rock potential of shallow water evaporitic settings; Journal
of Sedimentary Petrology, v. 56, p. 442454.
Wasson, R.J., Smith, G.I., and Agrawal, D.P.
1984: Late Quaternary sediments, minerals, and inferred geochemicalhistory of Didwana Lake, Thar Desert, India; Palaeogeography,
Palaeoclimatology,Palaeoecology, v. 46, p. 345-372.
Weare, J.H.
1987: Models of mineral solubility inconcentrated brines with application
to field obselvations; Reviews in Mineralogy, v. 17, p. 143-177.
Wood, J.R.
1975: Thermdynamics of brine-salt equilibria - I. the systems
NaCI-KCl-MgC12-CaC12-H20
and NaC1-MgS04-Hz0 at 2SC;
Geochimica et Cosmochimica Acta, v. 39, p. 1147-1 163.
Xia, J., Haskell, B.J., Engstrom, D.R., and Ito, E.
1997: Holocene climate reconstructions from tandem trace-element and
stable isotope composition of ostracodes from Coldwater Lake,
North Dakota, USA; Journal of Paleolimnology, v. 17, p. 85-100.

Late Holoceiie paleolimnology of


Killarney Lake, Manitoba
Kelly-Anne ~ichrnond'and L. Gordon ~ o l d s b o r o u ~ h '
Richmond, K.-A., and Goldsborough, L.G., 1999: Late Holocene paleolimnology of Killarney
Lake, Manitoba; in Holocene Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Contextfor Evaluating the Impacts of Climate Change on the Southern Canadian
Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological Survey of Canada, Bulletin 534,
p. 111-123.

Abstract: Analyses of organic matter, total phosphorus, pigments (chlorophyUs, total carotenoids,
myxoxanthophyll, oscillaxanthin), and diatoms were used to infer late Holocene trends in primary production for Killarney Lake, southwestern Manitoba. At 4700 BP the lake was relatively shallow and unproductive. Water levels and production then rose, likely due to a cooler and wetter climate. A deep, unproductive
lake existed by about 2100 BP. Water levels decreased and production increased about 1200 BP. Water levels increased again about 500 BP but didnot reach the level of the previous deep-water stand. Production has
increased during the last 100 years, possibly due to anthropogenic influences in the watershed. Collectively,
these analyses indicate that Killarney Lake has been mesotrophic to eutrophic for at least the last five millennia. The similarity of its developmental sequence to those of other prairie lakes, especially in eastern
Saskatchewan, indicates a close linkage between lake-water level, production, and regional climate.

RCsumC : Des analyses de la matibre organique, du phosphore total, des pigments (chlorophylles,
carotbno'ides totaux, myxoxanthophylle, oscillaxanthine) et des diatom6es ont CtC utilisCes pour dCduire
1'Cvolution au cours de 1'Holockne supbrieur de la productivit6 primaire au lac Killarney, dans le sud-ouest
du Manitoba. A 4 700 BP, le lac 6tait relativement peu profond et improductif. Le niveau de I'eau et la
productivit6 se sont par la suite accrus, vraisemblablement en rais?n d'un climat plus frais et plus humide.
Un lac profond et irnproductif s'ttait form6 I? environ 2 100 BP. A 1 200 BP environ, le niveau de I'eau a
diminu6 et la productivitb s'est accrue. Le niveau de l'eau a de nouveau augment6 Benviron 500 BP mais n'a
pas atteint le niveau d'eau profonde antkrieur. Depuis un sibcle la productivit6 s'est accrue, peut-Ctre en raison d'influences anthropogbnes dans le bassin versant. L'ensemble de ces analyses indique que le lac
Killarney varie de mCsotrophe 21 eutrophe au moins depuis cinq millCnaires. La similitude de son Cvolution
avec celle d'autres lacs des Prairies, surtout dans l'est de la Saskatchewan, tkmoigne de liens 6troits entre
niveau de l'eau, productivit6 et climat rbgional.

University Field Station (Delta Marsh) and Department of Botany, University of Manitoba, Winnipeg, Manitoba
R3T 2N2

GSC Bulletin 534

INTRODUCTION
Lacustrine sediments preserve a detailed record of biotic
changes occurring in a lake basin and its watershed.
Autochthonous biotic materials (algae, plants, invertebrates,
fish), as well as allocthonous inputs from the watershed
(particulates, litter), are deposited on the lake bottom.
Depending on their chemical composition and the environmental conditions at the sediment surface, these materials can
persist, preserving a record of conditions in and around the
lake at the time of deposition. Analysis of sediment cores for
constituent chemical, physical, and biological parameters is a
powerful technique for studying changes in lake production
due to changes in regional and local climate, hydrological
patterns, nutrient loading, and other environmental factors.
records in the
Canadian Prairies
are an underutilized archive of information on past hydrological and climate change in this climatically sensitive region
(Vance and
1994;
et
1997)' In particular,
there are few hindcasts of postglacial 'Iimate
in western
Manitoba' Most
using
pollen as a proxy for plant
communities occurring around the lake, was done by Ritchie
during the 1960s in the Riding Mountain area (Ritchie, 1969),

at Sewell Lake southeast of Brandon (Ritchie, 1976), and in


the Tiger Hills of south-central Manitoba (Ritchie and
Lichti-Federovich, 1968) (Fig. 1). Paleoenvironmental
reconstructions from these sites are in general agreement. A
boreal forest at about 10 000 BP gave way to grassland,
presumably in response to warmer, drier conditions. The
modern mixed-wood forest was established about 2500 BP,
as the climate became cooler and wetter (Ritchie, 1969).
Regrettably, much of the data from these studies was
published only in summary form, and no sites from the
mixed-grass aspen parkland of southwestern Manitoba have
been examined. In addition, there were no data on historical
changes in lake production, water quality, and water level, all
factors that might be affected by changes in climate.
The objective of this project was to use changes in the concentrations of organic matter, phosphorus, diatom valves, and
fossil pigments, measured in a core of lacustrine sediments, to
infer late Holocene trends in primary production for a moderately deep freshwater lake in
Manitoba. The
results provide a basis for evaluating modern waterquality
concerns and a benchmark against which anthropogeniccontamination by domestic and agriculturalwastes may be evaluated. It should also provide a baseline for evaluating the
impact of climate change, particularly in the context of other
lakes in more arid regions-of the prairie Provinces.

Town
of
Killarnav

Figure 1. Location of coring sites in Killarney Lake, southwestern Manitoba. Lake


bathyrnetry in metres provided by the Manitoba Department of Natural Resources
(date of determination unknown). Core K2 was collected at site I and core K3 was collected at site 2. Locations of other prairie lakes mentioned in the text: I ) Manitoba,
2 ) Sewell, 3 ) Riding Mountain unnamed lakes, 4 ) Kenosee, 5) Ceylon, 6) Waldsea,
7)Harris, and 8) Chappice.

K.-A. Richmond and L.G. Goldsborough

SITE DESCRIPTION

MATERIALS AND METHODS

Gllarney Lake (lat. 49011fN, long. 9g042'W, elev. 490 m


a.s.1.) is situated in the Pembina River Plain of southwestern
Manitoba (Fig. I), with Turtle Mountain upland to the south
and the Pembina Channel to the north. It occupies an elongate
shallow basin (maximum depth 7.0 m, surface area 163.6 ha)
and is bordered by the town of Killarney, established in the
1880s. The lake has no permanent overland inflow or outflow, although spring floodwater from the adjacent Long
River is diverted into the lake via a channel constructed in
1956. The amount of lake inflow and outflow is small compared to the total volume that bypasses it in the Long River.

Coring

Killarney Lake occurs in the aspen parkland ecoregion of


the prairie ecozone (Ecological Stratification Working
Group, 1995). Native vegetation in the watershed consists of
mixed short- and tall-grass species with associated herbs.
Clumps of bur oak (Quercus macrocarpa Michx.) occur
along the north shoreline, and trembling aspen (Populus
tremuloides Michx.), green ash (Fraxinus pennsylvanica
Marsh.), and willows (Salix spp.) occur on the south shore.
Cultivated fields contact the shoreline in many places or are
separated from it only by a narrow buffer of trees. The emergent flora of the littoral zone consist of cattail (Typha sp.),
bulrush (Scirpus spp.), sedges (Cyperaceae), grasses
(Gramineae), and duckweed (Lemna minor L.).
Mean annual temperature (1961-1990 normals) at
Killarney Lake is +2.4"C. Total annual precipitation is
approximately 550 mm, 75% of which occurs as rain, and the
annual precipitation deficit is about 200 rnrn (Ferguson et al.
1970). Lake water is alkaline (65 m g . ~ -Ca,
l 46 m g . ~ -Mg,
'
47 mg.Lm1Na, 14 m g . ~ - lK, specific conductance
860 pScm-l; Manitoba Department of Environment, unpub.
spring data, 1991). Surface water sampling throughout the
ice-free period of 1991 (Table 1) indicated that the lake is
eutrophic (20.9 pg.L-l total chlorophyll) and phosphorus limited (molar N:P approx. 36), supporting periodic blooms of
cyanobacteria (Aphanizomenon flos-aquae (L.) Ralfs,
Microcystis aeruginosa Kiitz., and others).

Table 1. Chemistry of surface water samples (15 cm


depth) collected at six sites in Killarney Lake between April
and October 1991 (L.G. Goldsborough, unpub. data,
1991).
Parameter

Mean

Range

Secchi disk depth (m)

1.4

0.9-2.3

Dissolved oxygen ( m g t ' )

6.8-1 3.4

PH
Total alkalinity (rng.LS')

8.6
172

8.2-9.1
159-203

Total Kjeldahl nitrogen (mgU1)

1.28

0.82-1.61

Ammonia-N (mg.L-')

0.1

0.03-0.25

(Nitrate+nitrite)-N (mgC1)

0.06

0.01-0.41

Total phosphorus (mg.L")

0.09

0.05-0.16

Total dissolved phosphorus (rng.L")

0.06

0.03-0.12

Total chlorophyll (pg.L-')

21

5-48

Two sediment cores were collected using a 7.6 cm diameter


percussion corer (Reasoner, 1986) from two sites in the deepest part of the lake basin on March 15, 1992 (Fig. 1). Water
depths at coring sites 1 and 2 were 6.9 and 6.2 m, respectively.
Core K2 from site 1 was 270 cm long and core K3 from site 2
was 280 cm long. Following retrieval, each core barrel was
cut at the sedimentlwater interface and sealed at both ends.
The cores were stored at 4C until analyzed.
Subsampling was performed within a 1 cm slice at 5 cm
intervals. For each slice, analyses of bulk composition, chlorophyll, and diatoms were performed on 1.9 g samples of wet
sediment, and 10 g of wet sediment were used for carotenoid
pigment analyses. Subsampling was restricted to the centre of
each core to minimize contamination from material transferred along the core barrel during coring.

Bulk composition
Water content, organic matter, and total carbonates were
determined by loss-on-ignition (Dean, 1974). Water content
(% wet weight) was measured as weight loss after drying at
105C for 24 hours, organic matter (% dry weight) after ignition at 550C for 1 hour, and total carbonates (% dry weight)
after further ignition at 950C for 3 hours.
The phosphorus concentration in the water column of
lakes often correlates closely with extant primary production
(Wetzel, 1983); therefore, since sediments are effective P
sinks, levels of P in the sediment may relate to past production
in the lake. An ignition method was used to determine the
total P of sampled sediments (Andersen, 1976). Total
extractable P was determined by combusting samples at
550C for one hour to mineralize organic matter, then boiling
the residue in 1 N HC1 to convert polyphosphates to
orthophosphate. The orthophosphate concentration (pg-g-'
dry weight) was determined using the molybdenum blue
method (Stainton et al., 1977).

Pigments
Preserved sedimentary pigments are used as indicators of past
lake production on the assumption that their abundance was
determined, in part, by the abundance of live algae and plants
at the time of their deposition (Leavitt, 1993). Chlorophyll
pigments in wet sediment samples were extracted in 90%
methanol for 24 hours in the dark. Solvent absorption at
665 nm and 750 nm, before and after acidification, was measured using a spectrophotometer. Concentrations of chlorophyll and pheophytin were calculated using the formulae of
Marker et al. (1980) and summed as total chlorophyll (pg.g-l
dry weight).
Bacteriochlorophyll a is produced by green and purple
bacteria that use H2S as an electron donor in anoxygenic photosynthesis (Pfennig, 1978), so its presence can indicate
anaerobic in situ conditions, even in relatively shallow water

GSC Bulletin 534

(e.g. Goldsborough and Brown, 1991). It was detected in core


samples by the appearance of near-infrared (750-770 nm)
absorption peaks in methanol extracts. Total carotenoid pigments, and the cyanobacterial pigments myxoxanthophyll
and oscillaxanthin, were quantified in the cores following the
methods of Sanger and Gorham (1972) and Swain (1985).
Pigment concentrations (pg.g-l dry weight) were calculated
as in Swain (1985). Planktonic cyanobacteria typically thrive
under nutrient-rich conditions (Wetzel, 1983), so the presence of myxoxanthophyll and oscillaxanthin may indicate the
development and magnitude of lake eutrophy (Swain, 1985;
Hickman and Schweger, 1991).

Diatoms
Diatoms occur across a wide range of environmental conditions and the niches of individual species are well known (e.g.
Wilson et al., 1997). Siliceous diatom valves can persist in
sediments and are readily identified to species, thereby providing a metric of total production as well as detailed ecological information (Dixit et al., 1992). Diatom valves in K2 core
samples were separated from the sediment matrix using a
series of water washes. Distilled water was added to a wet
sediment sample and the sample was sonicated for about
15 minutes to facilitate breakup of sediment aggregates. A
subsample was then washed three times with 1% calgonT"in
distilled water, followed by at least six distilled water washes.
The settling time between washes was 3 hours, as determined
using formulae in Folk (1965). Microscopic examination of
the discarded supernatant showed that this time was adequate
to permit quantitative isolation of diatoms from clay-size particles in suspension. A measured volume of a well agitated
suspension of diatom valves was deposited onto glass
coverslips in settling trays (Battarbee, 1973). The dry slips
were combusted at 550C for six minutes to remove organic
matter, then mounted on slides using NaphraxTMmounting
medium (Northern Biological Supply, England). A minimum
of 500 valves were identified and enumerated from each sample at lOOOx magnification using phase-contrast optics. Diatom concentration
dry weight) was calculated
based on the known sediment mass on the slide and the proportion of coverslip area examined during counting.
Species identification was based on Patrick and Reimer
(1966,1975), Germain (l981), Hustedt (1985), Krarnmer and
Lange-Bertalot (1986, 1988, 1991a, b), Lipsey (1987), and
Kling and H&ansson (1988). Common diatom taxa were
Table 2. Conventional radiocarbon dates on bulk sediments from Killarney Lake cores. Analyses performed at
Alberta Environmental Centre-Vegreville (AECV).
Depth range

number
AECv

(years
DateBP)

defined as those occurring in at least two samples at an abundance greater than or equal to 1% of the total. Taxonomically
consistent assemblages were identified using cluster analysis
(Euclidean distance and sum-of-squares distance method) on
log-transformed diatom concentration data (Wolin, 1996).
Diatom 'groups' were defined by occurrences of specific
clusters along the core length. Species niche preferences
(planktonic or benthic) were assigned using ecological information from Patrick and Reimer (1966, 1975), Lowe (1974),
Beaver (l981), andL.G. Goldsborough (unpub. data, 1999).

Sediment chronology
No identifiable macrofossils were found in either core that
could be dated via accelerator mass spectrometry (AMS).
Bulk sediment samples were collected at three positions in
each of cores K2 and K3 for conventional radiocarbon dating
(Table 2). The samples were air dried, wrapped in aluminum
foil, and stored at 4OC prior to analysis. Analyses were performed at the Environmental Isotopes Lab (Alberta Environmental Centre-Vegreville). These dates were used to interpolate sediment ages for other burial depths. Several lines of
evidence suggest that these dates were not erroneously old as
a result of contamination. First, a 420 cm long core collected
at site 1 (Fig. 1) in March 1993 (Vance and Last, 1994)
yielded Scirpus sp. seeds in two strata that were dated by
AMS at 9180 BP (345-350 cm) and 6090 BP (292-300 cm).
Sedimentation rates based on these values gave markedly
older, not younger, dates for our sampled strata. Second, our
dates from both cores, plotted as a function of sediment depth,
extrapolated close to the origin (0 years BP). Third, the water
of Killarney Lake is fresher than that of many lakes on the
Prairies, so the significance of old carbon derived from leaching of calcareous glacial materials (the 'hardwater effect';
Vance et al. (1993)) should be less. Finally, we saw no evidence of coal, a common constituent of sediments from
Alberta lakes (Hickman and Schweger, 1993), during microscopic examination of our intact, wet sediments.

RESULTS
The cores were nonlaminated dark olive grey (5Y 312;
MunsellTMColour Chart) gyttjas, although the bottom 15 cm
of core K2 was coarse sand and gravel. Radiocarbon dating of
the organic stratum immediately above the sand indicated
that the Killarney Lake basin is at least 4670 14cyears old.
Although the core lengths were similar, the K3 record was of
shorter duration, representing about 3200 14cyears. The
dates obtained for the K2 and K3 cores indicated that the average sedimentation rates were about 0.59 and 0.75 rnm.a-',
respectively.

Bulk composition and pigments


Variations in water and carbonate content along the cores
were small (Fig. 2,3). Water content in K2 (mean 76%) was
low in the sand layer at 270 cm but increased to about 85% at
the surface. There was no corresponding dry layer at the
bottom of K3. Otherwise, water content was similar to K2

K.-A. Richmond and L.G. Goldsborough

Total
Myxoxanthophyll I4cage
carotenoids Oscillaxanthin
(c~s.g-')
(IJg.g-')
(years BPI
(1.19.9-I)
0
20 0
20
40 0
20
40
1

490 & 70

2440 + 70

'

Figure 2. Changes in water content, total carbonates, organic matter, total phosphorus, total chlorophyll,
total carotenoid pigments, oscillaxanthin, and myxoxanthophyll along the K2 core from Killarney Lake.

Water
(%I

Total
Carbonates
(%)

Organic
matter
(%I

Total
Phosphorus
(~g.g"I

Figure 3. Changes in water content, total carbonates, organic matter, total phosphorus, total chlorophyll,
total carotenoid pigments, oscillaxanthin, and myxoxanthophyll along the K3 core from Killarney Lake.

4670

* 80

GSC Bulletin 534

(mean 82%). Total carbonates averaged 10-1 1% over the


length of both cores, except for a low value (6%) at 270 cm in
K2.
Concentrations of organic matter, total phosphorus, total
chlorophyll, and total carotenoids were more variable along
the core lengths than water and carbonate content, but depth
ranges for high and low values of all parameters generally
coincided in both cores (Fig. 2,3). Concentrations were low
in the bottom sand layer of K2, and increased gradually to
maxima between about 3000 and 1800 BP (1 10-175 cm in
K2,150-250 cm in K3) and between about 1500 and 300 BP
(20-80 cm in K2,20-120 cm in K3). Noticeably low values
occurred about 1500-1800 BP (80-110 cm in K2,
125-135 cm in K3) and in the uppermost sediments (top
20 cm, <300 BP).
Concentrations of fossil cyanobacterial pigments
(oscillaxanthin and myxoxanthophyll) fluctuated irregularly
throughout the cores, generally following changes in organic
matter, phosphorus, chlorophyll, and carotenoid concentration (Fig. 2, 3). Values were consistently low in the uppermost 25 cm of both cores. Peaks occurred about 400-800 BP
(25-60 cm), about 2100BP (1 10-130 cm in K2,160-170 cm
in K3), about 2700 BP (155 cm in K2, 190-220 cm in K3),
about 3200 BP (170-200 cm in K2,250-275 cm in K3), and
about 4200 BP (250 cm in K2).
Evidence of bacteriochlorophyll in the cores was equivocal. There were no near-infrared absorption peaks in solvent
extracts of K2 samples, but absorption of K3 samples
between 200 and 235 cm (ca. 3100-2700 BP) was about two
times higher than in other samples.

Total
diatoms
'
O va1ves.g.')
(10

Relative
abundance
(%)

Planktonic:
benthlc
ratio

ShannonWiener
diversity

Diatoms
Diatom valves were found at all levels of the K2 core (Fig. 4)
with a mean abundance of 3.70 x lo9 va1ves.g-l. Diatoms
were least abundant at the base of the core (270 and 190 cm;
ca. 4700-3200 BP). Elevated concentrations occurred
between 75 and 120 cm (ca. 1300-2000 BP) and in the top
10 cm (ca. 100BP to present). A total of 108 diatom taxa were
identified in the core, 82 of which were characteristic of benthic habitats, 18 of planktonic habitats, and 8 of either habitat.
However, planktonic diatoms were numerically dominant,
contributing greater than 80% of total valves in most samples
(Fig. 4). Planktonic taxa were particularly notable between
120 and 75 cm (ca. 2000-1300 BP), where the ratio of planktonic to benthic valves exceeded 30; otherwise, the ratio was
generally about 9. The Shannon-Wiener diversity of diatoms
ranged from 0.8 to 3.9 with a mean of 2.1.
Cluster analysis identified five distinct groups of diatom
species composition (Fig. 5). Group A contained samples
from the uppermost 15 cm and was characterized by the
planktonic taxa Asterionella fomzosa Hass. and Fragilaria
crotonensis Kitt. Group B contained samples from
270-220 cm (ca. 4700-3700 BP) and from 25-20 cm
(ca. 400-300 BP), both of which were defined by epipelic
(sediment-associated) taxa (e.g. Amphora ovalis Kutz. and
Fragilaria brevistriata Grun.). Group C included samples
from 105-80 cm (ca. 1800-1400 BP) that were dominated by
planktonic diatoms, such as Cyclotella bodanica Eul. and
Aulacoseira spp. (A. granulata (Ehr.) Simons., A. ambigua
(Grun.) Simons., and A, subarctica (0.Mull.) Haworth, the
former two of which were grouped because they could not be
distinguished reliably with light microscopy). Group D
contained samples from 160-135 cm (ca. 2800-2300 BP) and
from 35-30 cm (ca. 600-500 BP), in which both benthic and
planktonic taxa were abundant. Finally, group E was defined
''C age

(years
Bp)

Figure 4.
Total concentration, proportionate abundance of planktonic and benthic taxa, and
Shannon- Wiener (base e) diversity of diatoms
in the K2 corefronz Killarney Lake.

K.-A.

by samples in which benthic taxa such as Fragilaria


brevistriata and F. construens (Ehr.) Grun. were dominant,
including those from 215-165 cm (ca. 3700-2800 BP),
130-110 cm (ca. 2200-1900 BP), and 75-40 cm
(ca. 1300-700 BP).
Sum-of-squares distance

Richmond and L.G. Goldsborough

Planktonic centric diatoms were abundant throughout the


core (Fig. 6). Of these, Aulacoseira spp. were most frequent,
occurring in nearly all samples and with peak abundance
between 130 and 75 cm (ca. 2200-1300 BP). Other common
species included Stephanodiscus niagarae Ehr., Cyclotella
bodanica (most abundant at 85 cm but absent from the uppermost 7 0 cm), and several small Cyclotella and
Stephanodiscus taxa ( C , meneghiniana Kiitz., C.
michiganiana Skv., S. agassizensis H i k . & Kling,
S, hantzschii Grun., and S. minutulus (Kutz.) C1. &Moll.) that
we could not differentiate consistently by light microscopy.
This latter group was most abundant at 150 cm (ca. 2600 BP)
and in the top 10 cm (ca. 100 BP to present).
Among pennate diatoms, three small Fragilaria taxa were
common. F. brevistriata was abundant in most portions of the
core, particularly near 70 cm (ca. 1200 BP) and 120 cm (ca.
2000 BP). F. construens was also abundant in these strata, as
well as at 185 cm (ca. 3200 BP). F. pinnata Ehr. was less
abundant than the other two taxa but its stratigraphic distribution was similar to that of F, construens. These last two taxa
were rare in the bottom 40 cm (ca. 4700-4000 BP) and top
25 cm (ca. 400 BP to present) of the core.
Several pennate taxa were only abundant below 140 cm
(ca. 2400 BP), including Cymbella cymbi$ormis (Ag. Kutz.)
V . Heurck and C. hustedtii Krasske. Amphora pediculus
Kiitz., Cymbella muelleri Hust., Epithemia smithii
Carmthers, Navicula cuspidata Kiitz., and N. oblonga Kutz.
were abundant between 225 and 130 cm (ca. 3900-2200 BP)
and between 100 and 25 cm (ca. 1700-400 BP). These taxa
were absent from the top 25 cm (ca. 400 BP to present) of the
core. The stratigraphic distributions of Amphora ovalis,
Cocconeis placentula Ehr., Gomphonema subclavatum
(Grun.) Grun., and Synedra ulna (Nitz.) Ehr. were similar and
exhibited maximum abundance between 150 and 160 cm
(ca. 2800-2600 BP).
A few taxa, most notably Asterionella formosa,
Fragilaria crotonensis, and Nitzschia sigmoidea (Ehr.) W .
Sm., were found exclusively in samples near the top of the
core. The former two diatoms were absent from all samples
below 30 cm (ca. 500 BP).

INTERPRETATION

Figure 5. Cluster analysis (sum of square Euclidean distance


method) of diatom composition in samples collected at 5 cm
intervalsfrom the K2 corefrom Killarney Lake. Shaded boxes
denote sample groups having similar diatom composition
(see text for explanation).

Sediment cores from Killarney Lake include about 4700


years of deposition, starting in the warm and dry middle
Holocene. However, they do not constitute the entire
postglacial record because a 4.2 m long core with a basal date
of at least 9200 years was retrieved at our site using a
vibracorer (Vance and Last, 1994). The coarse sand and
gravel at a sediment depth of about 2.8 m, recovered partially
in the lowermost 15 cm of the K2 core, apparently prevented
further penetration of the percussion corer. We speculate that
this layer was deposited during a period of watershed instability and low lake level. Prairie lakes containing mineral-rich
sediments commonly contain desiccated and cemented horizons (Vance et al., 1992),which may correspond to periods of

GSC Bulletin 534

lake drawdown and pedogenesis (e.g. Lake Manitoba; Teller


and Last (1982))due to dry or extremely low water conditions
between ca. 9500 and 4500 BP.
The sand layer is absent from the K3 core because it represents a shorter chronological sequence, despite being of
similar length to K2, a result that we believe reflects differences in sedimentation rate between the two coring sites and
differential compaction during core collection. Results from
the two cores agree closely when strata are matched on the
basis of date rather than depth, indicating that the cores are
representative of lake-wide changes in autochthonous
production.
Changes in the paleoproduction of Killarney Lake have
been conservative. The diatom flora in the lake, for example,
have not changed substantially, being generally typical of
alkaline, nutrient-replete water. Changes in species composition appear to be related to fluctuating water levels, as a determinant of benthic and planktonic habitat availability, rather
than changing nutrient status. There was no evidence of
profound paleosalinity changes. Anaerobiosis, indicated by

the presence of bacteriochlorophyll, was limited, implying


that either 1) the lake remained sufficiently deep that anoxic
conditions did not develop in hypolimnetic waters, or 2) the
lake was not productive to the extent that massive algal
decomposition and self-shading would lead to anoxia.
For the purpose of interpretation, we consider the history
of Killarney Lake over the past 4700 years to consist of five
phases (Fig. 7). We base this chronology mostly on the K2
core because it represents the longest depositional record.

Phase 1 - infilling of a dry lake (ca. 4700-3000 BP)


Killarney Lake was shallow and unproductive during this
period, possibly recovering from a period of being dry or
nearly so. This is indicated by the sand and gravel at the base
of the K2 core, the abundance of benthic diatom taxa, and the
relative scarcity of centric diatoms, typical of planktonic conditions in a deeper lake (Fig. 6, group B,). Large quantities of
myxoxanthophyll and oscillaxanthin may be due to
cyanobacterial crusts (e.g. Oscillatoria spp.) on the bottom of
this shallow system. Such algae are common in the modern

c
Benthic

Concentration (10' valves.g-')

Figure 6. Concentrationof the most common diatom taxa enumerated in the K2 corefrom Killarney
Lake. Diatom zones were assignedfrom individual occurrences offive core sample groups identified
by cluster analysis of species abundance data (Fig. 5).

K.-A. Richmond and L.G. Goldsborough

epipelon of prairie wetlands (L.G. Goldsborough, unpub.


observation, 1997). Over a period of about 1000 years, the
water level probably increased gradually, perhaps due to
increased precipitation, causing production to decrease as
cyanobacterial crusts waned and were replaced by epipelic
diatoms such as Fragilaria spp. (Fig. 6, group El). Invasion
of the developing littoral zone by submersed macrophytes
and their associated epiphytic (attached) algae, along with
few planktonic algae in the limited pelagic zone, would lead
to a gradual increase in lake production.

Phase 2 - littoral development in a shallow lake


(ca. 3000-1800 BP)
Water levels and production began to rise in Killarney Lake,
probably in response to a cooler and wetter climate, about
3000 BP. The trend commenced in the late stages of Phase 1
and continued to about 1800 BP, with the increasing abundance of planktonic diatoms such as Aulacoseira spp, suggesting a deepening water column in a eutrophic lake, abundant macrophytes, and epiphytic algae in an expansive,

development o f a deepening lake


phase 3
(ca. 1800-1500 BP)
Production in Killarney Lake, as inferred from low levels of
organic matter and fossil pigments, was low during a wetter
period between about 1800 and 1500 BP (Fig. 7). The
increase in planktonic diatoms (Aulacoseira spp.) and a

Planktonic

Concentration (10'

Figure 6. (cont.)

highly productive littoral zone. Organic matter, pigments,


and total diatoms in the sediments from this period were high,
and characteristic epiphytic diatom genera, including
Amphora, Cocconeis, Cymbella, and Epithemia (Fig. 6,
group Dl), were numerous. Planktonic cyanobacterial
blooms, indicated by high myxoxanthophyll and oscillaxanthin levels, probably occurred frequently. Planktonic
diatoms began to displace benthic forms late in this period
probably indicative of ongoing pelagic
(Fig. 6, group b),
development of the lake to its deepest point by about 2100
BP. Production at this time was low, compared to a peak
about 2500 BP that corresponded to a period of abundant
epiphytic diatoms and cyanobacterial pigments.

GSC Bulletin 534

1490 f 70

1 2440 t 70

low but
increasing

low but
increasing

planktonic
taxa high

maximum
peak

peak

maximum
peak

benthic
taxa high

low

low

low but
increasing

peak

high

low

high

low

low

higher
water level

Fragilaria spp.

low
water level,
expansion of
littoral zone

AulacOseira 'pp.
max. planktonic
,in, benthic Stephanodiscusniagarae
Cyclotella bodanica
planktonic
taxa increase
benthic
taxa increase

low

Asterionella formosa
Fragilaria crotonensis

benthic
diatoms

Aulacoseira spp.
Amphora spp.
Fragilaria spp.

1 4670

* 80

low

high

....

I.......

.,..-.
*..

i,+
higher
water level
expansion of
littoral zone

....
Fragilaria spp.

low total
diatoms
high

-.a*-

recovery
from dry
or nearly

d'Y

cond~t~on

high

.a

Figure 7. Summary of inferred changes in the production ofKillarney Lake during the late Holocene, based
largely on results from the K2 core.

corresponding decline in benthic and epiphytic diatoms


(Fig. 6, group C) indicates a decrease in the proportion of
macrophytes and littoral habitat in general, probably due to
water levels being at or near their maximum for the entire
4700-year period. Myxoxanthophyll and oscillaxanthin levels were generally low, indicating that cyanobacterial blooms
were infrequent in a lake that probably was, at most,
mesotrophic.

Phase 4 - lake becomes shallower


(ca. 1500-300 BP)
After the deep-water stand around 2100 BP, Killarney Lake
became progressively shallower and more productive. This
was indicated by fewer numbers of planktonic diatoms, the
return of benthic and epiphytic diatom taxa (Fig. 6, group E3)
and their associated macrophytes to an enlarged littoral zone,
and increasing levels of organic matter, phosphorus, and pigments. Conditions similar to those in Phase 2 prevailed and
the lake was probably eutrophic, possibly even hypertrophic,
with abundant cyanobacterial blooms in the water column, as
indicated by maximal levels of myxoxanthophyll and
oscillaxanthin. Peaks of organic matter, chlorophyll, and
carotenoids indicate that maximum production occurred
about 700-500 BP. The lake began to deepen again around
500 BP, as indicated by the resurgence of planktonic diatoms

(Fig. 6, group D2) but it probably did not attain the level
achieved during Phase 3. Conditions similar to those in the
present lake were established about 300 BP.

Phase 5 - the modern lake ( 4 0 0 BP)


Fewer benthic diatoms and greater dominance by
Aulacoseira spp. indicate that Killarney Lake has become
deeper over the past 300 years. The development of a diatom
assemblage with no analogue in the earlier record (group A;
defined largely by the occurrence of Asterionella formosa
and Fragilaria crotonensis) suggests that the lake has
responded to settlement and agricultural development in its
watershed. The ensuing nutrient inputs have caused the lake
to become more eutrophic, although low levels of
myxoxanthophyll and oscillaxanthin suggest that
cyanobacteria, while frequent in the modern phytoplankton,
are probably less abundant than in the past, especially during
a period about 500 BP.

DISCUSSION
Comparison to other prairie lakes
Inferred trends in the water level of Killarney Lake agree generally with those from other lakes in southern Manitoba,
Saskatchewan, and Alberta, suggesting the influence of a

K.-A. Richmond and L.G. Goldsborough

regional climatic signal. Several studies of prairie lakes indicate a distinctly warmer and drier climate during the
mid-Holocene. In Lake Manitoba, lithological, rnineralogical, and chemical analyses indicated a period of fluctuating
wet and dry conditions from about 92004500 BP (Teller and
Last, 1982). Similarly, changes in physical, mineralogical,
and paleobiological parameters in sediments from Waldsea
Lake in south-central Saskatchewan indicated that a shallow
hypersaline lake with extensive mud flats existed about
4000 BP (Last and Schweyen, 1985).Maximum aridity in the
Cypress Hills of southwestern Saskatchewan occurred
approximately 7700-5000 BP, based on the vegetation
record from Harris Lake (Sauchyn and Sauchyn, 1991). Also
in Saskatchewan, Ceylon lake experienced high-salinity conditions with episodes of desiccation from about 8000 to
6000 BP and continued low lake levels until about 4000 BP
(Last, 1990; Teller and Last, 1990). Chappice Lake in southeastern Alberta oscillated between relatively high-water
stands and desiccation from about 7300 to 6000 BP, while
from about 6000 to 4400 BP, lake levels were consistently
low and production was high (Vance et al., 1993). At Moon
Lake in south-central North Dakota, a mid-Holocene period
of high salinity from about 7300 to 4700 BP indicated a high
precipitation deficit (Laird et al., 1996). These data support
the inference of widespread dryness throughout much of the
Great Plains during the period from about 8000 to 4000 BP.
Like Killarney Lake, many sites on the Prairies indicate a
climatic change from a more arid mid-Holocene to one of
increasing moisture about 4000 BP, which continued until
about 2000 BP. For example, Waldsea and Ceylon lake
records show increasing lake levels about 4000 BP, culminating in deep-water stands by about 3000 and 2000 BP,
respectively (Last and Schweyen, 1985; Teller and Last,
1990).Between about 4400 and 2600 BP, lake level was more
stable but gradually rising and production was high at
Chappice Lake, while a large, relatively fresh lake existed
from about 2600 to 1000 BP (Vance et al., 1993).
Our inferred trends of water depth correspond remarkably
well to those made for Kenosee Lake in eastern
Saskatchewan (Fig. 1) by Vance et al. (1997). There, an early
shallow-lake phase from about 4000 to 3000 BP gave way to
deeper water between about 3000 and 2000 BP. A deep-water
period characterized the mostrecent 600 years. The similarity
of the records between the two lakes, over 200 km apart, and
especially the period of wetter conditions from about 3000 to
2000 BP, supports the theory of regional climate as a driving
force for water-level change, contrary to the contention that
this represented a period of greater aridity at Moon Lake in
central North Dakota (Laird et al., 1996). One notable way in
which the Kenosee and Killarney records differ, however, is
that Killarney diatoms showed no indication of salinity
changes during increased water level, unlike profound
changes in Kenosee ostracodes due to high salinity during
infilling.
Low water level in Killarney Lake at about 1000 BP was
evident at several other prairie sites, including Chappice Lake
(ca. 1000-600 BP; Vance et al. (1993)) and Waldsea Lake
(ca. 1000-700 BP; Last and Slezak (1986)). These shallow-water stands roughly correspond to the Medieval Warm

Period, about 950-750 BP (Laird et al., 1996). Evidence of a


high-water stand at Chappice Lake from about 600 to 100 BP
(Vance et al., 1993) suggests that wetter conditions existed
subsequent to the Medieval Warm Period.
Unlike many other prairie lakes, Killarney Lake appears
to have remained fresh and moderately deep during the late
Holocene. The numerical dominance by planktonic centrics
in its diatom flora is a clear indication that Killarney Lake was
consistently deeper than lakes such as Harris in southwestern
Saskatchewan (Fig. I), where benthic diatoms (notably
Fragilaria species that occurred sporadically in Killarney
Lake) were overwhelmingly dominant throughout a
9200-year record (Wilson et al., 1997).

Context for modern water quality


The results of this study provide a baseline for evaluating
recent concerns of residents in the Killarney Lake watershed
regarding the frequent occurrence of cyanobacterial blooms
and their undesirable effects on water quality, fisheries, and
recreation. Some believe that eutrophication of Killarney
Lake has been accelerated in the century since the town was
established, a conclusion that is supported by the appearance
in surficial sediments of diatoms such as Asterionella
fomosa and Fragilaria crotonensis that are indicative of cultural eutrophication. However, as has been found for other
eutrophic lakes on the Prairies (Hickman and Schweger,
1991; R. Hall and P.R. Leavitt, unpub. data, 1997), the presence of myxoxanthophyll and oscillaxanthin with depth indicates that cyanobacterial blooms are not a recent
phenomenon. Rather, they have probably occurred irregularly throughout the period represented by these cores. Recent
(last 100 years) production appears to be low in comparison
with two maxima between about 3000 and 1800 BP, and
between about 1500 and 300 BP.
In the context of the prehistoric levels of production in
Killarney Lake inferred here, it appears that anthropogenic
influences on the currently deep lake have been minor. A
notable exception is the dramatic accumulation of total
extractable copper in surface sediments as a result of
lake-wide applications, since at least the 1950s, of aqueous
copper sulphate for control of algal blooms, to the point
where its concentration is about 50 times higher than in any
other lake or dugout sediment collected in Manitoba
(L.G. Goldsborough, unpub. data, 1993). These additions are
occasionally successful at treating the result of nutrient inputs
from agricultural land, cattle feed lots, golf courses, and
domestic wastes from the watershed. Otherwise, our analyses
indicate that Killarney Lake has been naturally mesotrophic
to eutrophic, and at times hypertrophic with profuse
cyanobacteria, for at least 4700 years.

ACKNOWLEDGMENTS
We thank all those who helped with core collection and analyses, including Mike Forster, Scott Gadsby, Tom Henderson,
Debbie Hysop, Hugh Kiffen, Don Lemmen, Rhonda
McDougal, Ryan McGregor, Pauline Morton, Iain Pimlott,

GSC Bulletin 534

Bob Vance, and Maria Zbigniewicz. Hedy Kling assisted


with diatom identification and Norm Kenkel assisted with
cluster analyses. We gratefully acknowledge financial support from the Manitoba Environmental Innovations Fund, the
Town of Killarney, the Rural Municipality of Turtle Mountain, the Geological Survey of Canada Palliser Triangle
Global Change Program, the Brandon University Research
Committee, a Natural Sciences and Engineering Research
Council of Canada (NSERC) Research Grant to L.G.G., and a
University of Manitoba Fellowship and Manitoba Naturalists
Society Scholarship to K.A.R.

REFERENCES
Andersen, J.M.
1976: An ignition method for determination of total phosphorus in lake
sediments; Water Research, v. 10, p. 329-331.
Battarbee, R.W.
1973: A new method for the estimation of absolute microfossil numbers,
with reference especially to diatoms; Limnology and Oceanography, v. 18, p. 647-653.
Beaver, J.
1981: Apparent ecological characteristics of some common freshwater
diatoms; Ontario Ministry of the Environment, Water Quality
Branch, Toronto, Ontario, 517 p.
Dean, W.E.
1974: Determination of carbonate and organic matter in calcareous sediments and sedimentay rocks by loss on ignition: colnparison with
other methods; Journal of Sedimentary Petrology, v. 44,
p. 242-248.
Dixit, S.S., Smol, J.P., Kingson, J.C., and Charles, D.P.
1992: Diatoms: powerful indicators of environmental change; Environmental Science and Technology, v. 26, p. 23-33.
Ecological Stratification Working Group (ESWG)
1995: A National Ecological Framework for Canada; Agriculture and
Agri-Food Canada, Research Branch, Centre for Land and Biological Resources Research and Environment Canada, State of the Environment Directorate, Ecozone Analysis Branch, OttawaIHull,
125 p.
Ferguson, H.L., O'Neill, A.D.J., and Cork, H.F.
1970: Mean evaporation over Canada; Water Resources Research, v. 6,
p. 1618-1633.
Polk, R.L.
1965: Petrology of Sedimentary Rocks; Hemphill's, Austin, Texas, p. 40.
Germain, H.
1981: Flore des DiatomCes - DiatomophycCes; SociCtC Nouvelle des
Editions BoubCe, Paris, France, 444 p.
Goldsborough, L.G. and Brown, D.J.
1991: Periphyton production in a small, dystrophic pond on the Canadian
Precambrian Shield; Internationale Vereinigung fur Theoretische
und AngewandteLimnologieVerhanglungen, v. 24, p. 1497-1502.
Hickman, M. and Schweger, C.E.
1991: OscilJaxanthin and myxoxanthophyll in two cores from Lake
Wabamun, Alberta, danada; ~ o ; b a l of Paleolimnology, v. 5,
p. 127-137.
1993: Lateglacial-early ~ o l o c e n e ~ a l a e o s a l i nini tAlberta,
~
Canada-climate implications; Journal of Paleolimnology, v. 8, p. 149-161.
Hustedt, F.
1985: The Pennate Diatoms; Koeltz Scientific Books, Koenigstein, Germany, 918 p. (translated from "Die Kieselalgen, 2. Teil" by N.G.
Jensen)
Kling, H. and Hikansson, H.
1988: A light and electron microscope study of Cyclotella species
(Bacillariophyceae) from central and northern Canadian lakes; Diatom Research, v. 3, p. 55-82.
Krammer, K. and Lange-Bertalot, H.
1986: Bacillariophyceae. I. Teil: Naviculaceae; in SiiBwasserflora von
Mitteleuropa, Band 211, (ed.) H. Ettl, G. Gartner, J. Gerloff,
H. Heynig and D. Mollenhauer; Gustav Fischer Verlag, Stuttgart,
Germany, 876 p.

Krammer, K. and Lange-Bertalot, H. (cont.)


1988: Bacillariophyceae. 2. Teil: Bacillariaceae, Epithemiaceae,
Surirellaceae; in SiiBwasserflora von Mitteleuropa, Band Z2, (ed.)
H. Ettl, J. Gerloff, H. Heynig, and D. Mollenhauer; Gustav Fischer
Verlag, Stuttgart, Gennany, 61 1 p.
1991a: Bacillariophyceae. 3. Teil: Centrales, Fragilariaceae, Eunotiaceae;
in SiiBwasserflora von Mitteleuropa, Band 213, (ed.) H. Ettl,
G. Girtner, J. Gerloff, H. Heynig, and D. Mollenhauer; Gustav
Fischer Verlag, Stuttgart, Germany, 576 p.
1991b: Bacillariophyceae. 4. Teil: Achnanthaceae Kritische Erganzungen
zu Navicula (Lineolateae) und Gomphonema; in SiiBwasserflora
von Mitteleuropa, Band 214, (ed.) H. Ettl, G. Gartner, J. Gerloff,
H. Heynig, and D. Mollenhauer; Gustav Fischer Verlag, Stuttgart,
Germany, 437 p.
Laird, K.R., Fritz, C.S., Grimm, E.C., and Mueller, P.G.
1996: Century-scale paleoclimatic reconstruction from Moon Lake, a
closed-basin lake in the northern Great Plains; Li~nnology and
Oceanography, v. 41, p. 890-902.
Last, W.M.
1990: Paleochemistry and paleohydrology of Ceylon Lake, a
salt-dominated playa basin in the northern Great Plains, Canada;
Journal of Paleolimnology, v. 4, p. 219-238.
Last, W.M. and Schweyen, T.H.
1985: Late Holocene history of Waldsea Lake, Saskatchewan, Canada;
Quaternary Research, v. 24, p. 219-234.
Last, W.M. and Slezak, L.A.
1986: Paleohydrology, sedimentology, and geochemistry of two
meromictic saline lakes in southern Saskatchewan; GCographie
physique et Quaternaire, v. 60, p. 5-15.
Leavitt, P.R.
1993: A review of factors that regulate carotenoid and chlorophyll deposition and fossil pigment abundance; Journal of Paleolimnology, v. 9,
p. 109-127.
Lipsey, L.L., Jr.
1987: Freshwater diatoms (Bacillariophyceae) from the northeastern glacial lake district of Wisconsin. I. Attheya, Cyclotella, Melosiru,
Rhizosolenia, and Stephanodiscus (Order Centrales); Rhodora,
v. 89, p. 261-278.
Lowe, R.L.
1974: Environmental requirements and pollution tolerance of freshwater
diatoms; United States Environmental Protection Agency, Report
67014-74-005. Cincinnati, Ohio.
Marker, A.F.H., Crowther, C.A., and Gunn, R.J.M.
1980: Methanol and acetone as solvents for estimating chlorophyll a and
phaeopigments by spectrophotometry; Archiv fiir Hydrobiologie
Beihefte, v. 14, p. 52-69.
Patrick, R. and Reimer, C.W.
1966: The diatoms of the United States exclusive of Alaska and Hawaii,
Volume 1; The Academy of Natural Sciences of Philadelphia,
Philadelphia, Pennsylvania, Monograph 13,688 p.
1975: The diatoms of the United States exclusive of Alaska and Hawaii,
Volume 2, Part 1 ; The Academy of Natural Sciences of
Philadelphia, Philadelphia, Pennsylvania, Monograph 13, 213 p.
Pfennig, N.
1978: General physiology and ecology of photosynthetic bacteria; in The
Photosynthetic Bacteria, (ed.) R.K. Clayton and W.R. Sistrom;
Plenum Press, New York, p. 3-18.
Reasoner, M.
1986: An inexpensive, lighweight percussion core sampling system;
GCographie physique et Quaternaire, v. 40, p. 217-219.
Ritchie, J.C.
1969: Absolute pollen frequencies and carbon-14 age of a section of
Holocene lake sediment from the Riding Mountain area of
Manitoba; Canadian Journal of Botany, v. 47, p. 1345-1349.
1976: The late-Quaternary vegetational history of the Western Interior of
Canada; Canadian Journal of Botany, v. 54, p. 1793-1818.
Ritchie, J.C. and Lichti-Federovich, S.
1968: Holocene pollen assemblages from the Tiger Hills, Manitoba;
Canadian Journal of Earth Sciences, v. 5, p. 873-880.
Sanger, J.E. and Gorham, E.
1972: Stratigraphy of fossil pigments as a guide to the post-glacial history
of Kirchner Marsh, Minnesota; Limnology and Oceanography,
v. 17, p. 840-855.

K.-A. Richmond and L.G. Goldsborough


Sauchyn, M.A. and Sauchyn, D.J.
1991: A continuous record of Holocene pollen from Harris Lake, southwestern Saskatchewan, Canada; Palaeogeography,
Palaeoclimatology,Palaeoecology, v. 88, p. 13-23.
Stainton, M.P., Capel, M.J., and Armstrong, F.A.J.
1977: The chemical analysis of fresh water; Fisheries and Environment
Canada, Fisheries and Marine Service, Miscellaneous Special
Publication No. 25, 180 p. (second edition).
Swain, E.B.
1985: Measurement and interpretation of sedimentary pigments; Freshwater Biology, v. 15, p. 53-76.
Teller, J.T. and Last, W.M.
1982: Pedogenic zones in postglacial sediment of Lake Manitoba,
Canada; Earth SurfaceProcesses and Landforms,v. 7, p. 367-379.
1990: Paleohydrologicalindicators in playas and saltlakes, with examples
from Canada, Australia, and Africa; Palaeogeography,
Paleoclimatology,and Paleoecology, v. 76, p. 215-240.

Vance, R.E. and Last, W.M.


1994: Paleolimnology and global change on the southern Canadian
Prairies; in Current Research 1994-B; Geological Survey of
Canada, p. 49-58.
Vance, R.E., Clague, J.J., and Mathewes, R.W.
1993: Holocene paleohydrology of a hypersaline lake in southeastern
Alberta; Journal of Paleolimnology, v. 8, p. 103-120.
Vance, R.E., Last, W.M., and Smith, A.J.
1997: Hydrologic and climatic implications of a multidisciplinary study of
late Holocene sediment from Kenosee Lake, southeastern
Saskatchewan, Canada; Journal of Paleolimnology, v. 18,
p. 365-393.
Wetzel, R.G.
1983: Limnology; Saunders College Publishing, Toronto, Ontario, 767 p.
(second edition).
Wilson, S.E., Smol, J.P., and Sauchyn, D.J.
1997: A Holocene paleosalinity diatom reco~dfrom southwestern
Saskatchewan, Canada: Harris Lake revisited; Journal of
Paleolimnology, v. 17, p. 23-31.
Wolin, J.
1996: Late Holocene lake-level and lake development signals in Lower
Herring Lake, Michigan; Journal of Paleolimnology, v. 15,
p. 19-45.

Multiproxy record of prairie lake response to


climatic change and human activity,
Clearwater Lake, Saskatchewan
Peter R. ~eavitt',Rolf D. vinebrookel, Roland I. v all' Susan E. wilson3,
John P. srno13, Robert E. vance4, and William M. Last 3
Leavitt, P.R., Vinebrooke, R.D., Hall, R. I., Wilson, S.E., Smol, J. P., Vance, R.E., and Last, W.M.,
1999: Multiproxy record of prairie lake response to climate change and human activity,
Cleanvater Lake, Saskatchewan; in Holocene Climate and Environmental Change in the Palliser
Triangle: A Geoscientific Contextfor Evaluating the Impacts of Climate Change on the Southern
Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological Survey of Canada,
Bulletin 534, p. 125-138.

Abstract: Multiple paleolirnnological indicators (mineralogy, diatoms, pigments, and macrofossils)


were used to quantify the relative impacts of recent climatic change and terrestrial disturbance on
Clearwater Lake, a closed-basin lake in southern Saskatchewan. Four periods were identified: 1) initial conditions (ca. AD 1600-1920), characterized by low salinity, planktonic algal taxa, and variable primary production; 2) ca. 1920-1940, characterized by declining water levels, slightly elevated salinity,
predominantly benthic diatoms, and increased primary production; 3) ca. 1940-1970, characterized by
slightly increased water levels and declining concentrations of indicator algal pigments; and 4) 1970-1993,
characterized by a unique combination of abundant Chara, historical maxima of pigments from
chlorophytes and siliceous algae, and highly variable water levels, ionic composition, and diatom species
composition. The timing of these changes was more similar to patterns of land use than to those of climatic
change, suggesting that paleoclimatic reconstructions of recent events (e.g. global warming) must also consider the influence of land-use practices and local hydrology on lake characteristics.

Rksum6 : De nombreux indicateurs palColimnologiques (minkralogie, diatomees, pigments,


macrofossiles) ont CtC utilisks afin de quantifier les incidences relatives des changements climatiques et des
perturbations terrestres rCcents sur le lac Clearwater, lac de bassin fermC du sud de la Saskatchewan. Quatre
pkriodes ont Ctk identifikes : (1) conditions initiales (environ 1600-1920), pkriode caractkride par une faible salinitC, des algues planctoniques et une productivitC primaire variable; (2) environ 1920-1940, pkriode
caractCris6e par une diminution du niveau de l'eau, une lkgbre augmentation de la salinitk, des diatomCes
principalement benthiques et uneproductivitC primaire accrue; (3) environ 1940-1970, pCriode caractkride
par une lkgbre hausse du niveau de l'eau et une diminution de la concentration de pigments d'algues
indicateurs; et (4) 1970-1993, pCriode caractCrisCe par une combinaison unique de l'abondance de charas,
de maxima historiques des pigments de chlorophytes et d'algues siliceuses, et de la forte variabilitk du
niveau de l'eau, de la composition ionique et de la composition des espkces de diatomCes. La chronologie de
ces changements ressemble plus aux configurations d'utilisation des sols qu'h celle des changements
climatiques, ce qui donne ?i penser que les reconstitutions palCoclimatiques des CvCnements rkcents (p. ex. le
rCchauffement global) doivent Cgalement prendre en considCration l'influence des pratiques relatives ?i
l'utilisation des sols et de l'hydrologie locale sur les caractCristiques des lacs.
Limnology Laboratory, Department of Biology, University of Regina, Regina, Saskatchewan S4S 0A2
Department of Biology, University of Waterloo, Waterloo, Ontario N2L 3G1
Paleoecological Environmental Assessment and Research Laboratory, Department of Biology, Queen's University,
Kingston, Ontario K7L 3N6
Geological Survey of Canada, 601 Booth Street, Ottawa, Ontario K I A 0E8
Department of Geological Sciences, University of Manitoba, Winnipeg, Manitoba R3T 2N2

GSC Bulletin 534

INTRODUCTION
The continued anthropogenic production of greenhouse
gases is expected to promote global warming and severely
increase the aridity of central continental regions, including
the Palliser Triangle (El-Ashry and Gibbons, 1988; Gleick,
1989). The Palliser Triangle occupies much of the Great
Plains of western Canada (Fig. I), an important agricultural
area whose high moisture deficit (>400 rnm-a-l)makes it sensitive to climatic change (Lemmen et al., 1993; Last, 1994).
Long-term meteorological records show a distinct warming
trend in the Canadian Prairies during the past century, particularly during the 1980s (Gullet and Skinner, 1992; Evans and
Prepas, 1996). As a consequence, water levels have declined
more than 3 m in some prairie lakes since 1970 (Vance and
Last, 1994).

Lakes in the
can provide sensitive
long-term records of past climatic conditions that may prove
important for determining the effects of future climate change
(Vance, 1997). In particular, the balance between precipitation and evaporation can regulate water levels, salinity, and
brine composition of closed-basin lakes (Wiche, 1986;Vance
and Last. 1994; Lent et al., 1995; Fritz. 19961, which in turn
affects lake production, community composition, and water
quality (Hammer, 1990; Evans and Prepas, 1996). For example, periods of increased salinity can cause taxonomic shifts
ihat-favour chlorophytes and- diatoms over filamentous

cyanobacteria (Marino et al., 1990; Evans and Prepas, 1996;


but see Hammer et al., 1983). In addition, climate-related
salinity fluctuations alter the species composition of diatom
assemblages in a predictable and quantifiable manner (Fritz
et al., 1991, 1993; Laird et al., 1996a, b). Because the sediments at the bottom of prairie lakes preserve a mineralogical
record of lake chemistry and the physical remains of aquatic
communities, it should be possible to obtain detailed records
of biotic responses to past climate change from an analysis of
sediments (e.g. Vance and Last, 1994; Fritz, 1996).
Standard methods for paleoclimatological research can be
problematic for use in dry prairie regions. Trees suitable for
tree-ring analyses are rare on agricultural plains (Fritz et al.,
1994); consequently, most tree-ring records come from the
eastern foothills of the Rocky Mountains (e.g. Case and
MacDonald, 1995) or from river valleys (e.g. Will, 1946;
Meko, 1992) that are under different climatic and hydrological con&ols than
prairies (Fritz et al., 1994). Furthermore, pollen studies have proven difficult because climatic
interpretations are obscured by the low taxonomic resolution
(family level only, e.g. 'grasses') of fossil records from grasslands (Vance and Mathewes, 1994; Fritz et al., 1994; Laird et
a,., 1996a), Finally, little is known of how the response of
individual lakes to climatic change may vary with hydrological setting and human land-use history (Adams et al., 1990;
Wilson et
1997).

Figure I. Map of the northern Great Plains showing the location of Clearwater Lake within the
Palliser Triangle. Dashed line marks limit of the Palliser Triangle; shaded area represents the Brown
Chernozemic Soil Zone.

126

P.R. Leavitt et al.

In this study, we used stratigraphic changes in the abundance of diatom valves, algal pigments, plant macrofossils,
and the physical and mineralogical properties of lake sediments to assess the effects of climate change and human
land-use on Clearwater Lake, Saskatchewan, during the past
400 years. Diatoms are abundant members of the algal flora
of inland saline lakes whose distribution is strongly related to
lake-water salinity (Fritz et al., 1993; Curnrning et al., 1995;
Gasse et al., 1995) and whose species composition allows
quantitative reconstruction of past lake salinity (Fritz et al.,
1993; Laird et al., 1996a, b). Similarly, carotenoids and
chlorophylls include compounds that allow quantification of
past changes in algal abundance and gross composition
(Leavitt, 1993; Leavitt and Findlay, 1994), whereas remains
of aquatic plants provide information on water depth, littoral
zone development, and gross floral community composition
(Vance et al., 1992,1997). Changes in sediment texture, bedding characteristics, mineralogy, and geochemical composition further facilitate inferences of past lake levels, watershed
weathering, and lake chemistry (Torgersen et al., 1986; Teller
and Last, 1990).

Study site
Clearwater Lake (lat. 50' 53' N, long. 107" 54' W, elev. 680
m a.s.1.) is centrally located within the Palliser Triangle, a dry
subhumid region of the Great Plains of western Canada
(Fig. 1). Annual precipitation averages 380 mm, whereas
about twice this amount is lost through evaporation
(Canadian National Committee for the International
Hydrologic Decade, 1978).The region experiences long winters with mean annual temperatures of 1.5"C and average
January temperatures of -15C. Lakes are normally ice
covered from November to April. Summers are short and
warm with average daily July temperatures of 19" C. Aerial
photographs show that the level of Clearwater Lake has fluctuated over the last few decades (Fig. 2). Changes in the presence of numerous ponds and the lake's northeastern bay
indicate that water levels were relatively high in 1956 and
during the 1970s, and low following droughts of the late
1930s and the 1980s.
The basin occupies a small catchment situated on The
Missouri Coteau, a relatively stony and infertile region of
hummocky moraine and poorly integrated drainage.
Table 1. Water chemistry of Clearwater Lake,
1938-1995. Ion concentrations in mmo1.L-' (with
m g . ~ - in
' parentheses).
Parameter

Average

Range

Ca2'
0.09-0.27 (4-1 1)
0.32 (8)
Mg2+
4.73-8.69 (1 15-21 1)
6.1 3 (149)
Na'
2.35-5.32 (54-1 22)
3.45 (79)
K'
0.07-0.71 (3-28)
0.51 (20)
HCO,+C0,28.81-1 3.94 (537-847)
10.34 (626)
1.39-3.25 (134-312)
2.00 (192)
SO;
CI0.54-1.16 (19-41)
0.76 (27)
TDS
923-1 565
1109
8.40-9.25
8.9
pH
TDS -total dissolved solids (in mg.L").

Clearwater Lake is small (0.5 km2), relatively shallow (maximum depth approx. 9 m), and has uncharacteristically low
salinity (approx. 1 g . ~ - 'total dissolved solids (TDS);
Table 1) given the high regional precipitation deficit (approx.
400 mm.a-I). Vance and Last (1994) suggested that
Clearwater Lake is influenced by fresh groundwater inputs,
although hydrological budgets are not available. The lake
lacks permanent stream inflow, but drains from a northeastern outlet and intermittent streams during periods of exceptionally high water. Lake water is alkaline (pH approx. 9) and
dominated by M ~ ~Na+,
+ , and HC03- ions. The lake is always
supersaturated with respect to most carbonate minerals and,
because of high MgICa ratios (>20), aragonite is precipitated
as the stable carbonate phase. Consequently, present-day sediments contain mainly endogenic aragonite, detrital carbonates, detrital siliciclastics, and organic matter (Table 2).
Small quantities of endogenic or authigenic protodolomite
(disordered, nonstoichiometric CaMg(C03)2) and
magnesian calcite also occur in modern sediments, and are
probably generated in response to short episodes of elevated
(>30) or reduced (c10) MgICa ratios (Miiller et al., 1972).
A small regional park and seasonal-use cottages occupy
more than two-thirds of the perimeter of the lake (Fig. 2). Cottage construction increased after 1940 and the lake remains a
popular recreation locale despite declining lake levels since
the 1970s. Other land-use features include a golf course to the
west and a refuse dump immediately south of the lake. Sparse

Table 2. Summary of average mineralogicaland grain size


characteristics of modern and late Holocene sediment in
Clearwater Lake. Ranges shown in parentheses.
Parameter
Organic matter (%)
Mean particle size (pm)
Standard deviation of mean size
Median particle size (pm)
Sand (%)
Silt (YO)
Clay (%)
Detrital fraction (%)
(Endogenic+authigenic)fraction (%)
Detrital components:
Total clay minerals (%)
Quartz (%)
Plagioclase (%)
K-feldspars (%)
Amphibole minerals (%)
Calcite (%)
Dolomite (%)
Endogenic and authiqenic components:
Mg-calcite (%)
Mol% MgCO, in Mg-calcite
Nonstoichiometric dolomite (%)
Mol% CaCO, in protodolomite
Magnesite (%)
Aragonite (%)
Gypsum (%)
Na-sulphate salts (%)
Pyrite (%)

n.o. - not present

Modern
15.0
39.7
32.9
32.4
13.6
80.8
5.6
79.4
20.6

1800-1 990
15.1 (10-20)
26.5 (6-60)
24.8 (4-36)
21.9 (4-55)
7.8 (0-32)
80.5 (48-92)
11.1 (1-42)
83.9 (37-89)
16.1(11-63)

51.9
12.0
6.1
n.p.
n.p.
4.2
6.5

49.4 (23-69)
10.4 (4-34)
7.8(0-20)
2.6 (0-34)
<0.1 (0-2)
5.5 (0-14)
8.0 (3-27)

2.0
5.3
0.8
63.5
n.p.
15.8
n.p.
n.p.
n.p.

0.5 (0-5)
10.3 (4-1 6)
1.2 (0-8)
67.4 (58-70)
<0.1 (0-3)
12.9 (0-29)
1.6 (0-22)
<0.1 (0-7)
<0.1 (0-3)

GSC Bulletin 534

historical data suggest that the lake has recently undergone


episodic deoxygenation, leading to occasional fish kills and
poor water quality (Vance and Last, 1996).

MATERIALS AND METHODS


Core collection and chronology (R.E.Vance)
Three short cores (57-88 cm) were collected from the southern basin of Clearwater Lake in February 1994 using a modified gravity corer (Turner, 1994). Sites were within 3 m of
each other. All cores were extruded from the 10 cm diameter

plexiglass coring tube at the lakeside. During extrusion, the


outer 1 cm layer of the core was removed to reduce sample
contamination from core smearing. One core (CWS1) was
sectioned into 1 cm intervals and dated using 2 1 0 ~analyses
b
and the constant rate of supply model (CRS; Appleby and
Oldfield (1978)). A second core (CWS2) was subsampled at
1 cm intervals for sediment mineralogy, particle-size distribution, fossil pigments, siliceous microfossils, and plant
macrofossils. The shortest of the cores (CWS3) was sectioned into 2 cm increments and used for additional
macrofossil analyses. Sediment accumulation rates were
assumed to be similar among cores.

Figure 2. Aerial photographs of Clearwater Lake showingfluctuations in water levels, 1939-1990. Photographs were taken: A) November 11,1939 (NAPLA6752-15); B) September 24,1956 (NAPLA15504-14);
C) May 20,1972 (NAPLA22663-196);D) May 26,1979 (NAPLA25126-59); and, E) June 19,1990 (NAPL
A27580-166). Scale bar represents 0.5 km in the 1979 photograph, but is only approximate in other
photographs.

P.R. Leavitt et al.

Lithostratigraphy (W .M. Last)


All mineralogical analyses were performed using standard
X-ray diffraction techniques (Klug and Alexander, 1974;
Hardy and Tucker, 1988) and a Philips PW1710 Automated
Powder Diffraction System. Mineral identification was aided
by the use of an automated peak-match computer program
(Marquart, 1986). Percentages of the various minerals were
estimated from diffractograms using a weighted peak-intensity
method modified after Schultz (1964). Nonstoichiometry of
the dolomite and calcite was determined by examining the
displacement of the dIo4 peak on a detailed (slow) X-ray
diffractogram, and was calculated according to Goldsmith
and Graf (1958). Organic matter and total carbonate mineral
content of the sediment were determined by weight loss on
heating to 500" C and 1000C, respectively (Dean, 1974).
Particle-size spectra for each sample were determined using
an automated laser-optical particle-size analyzer (GALAITM
CIS-1; Aharonson et al., 1986) after removal of organic
matter. Particle-size statistics were calculated according to
Allen (1981). Mineralogy, organic matter content, and
texture of sediments from CWSl were determined at 1 cm
intervals.

Diatoms (S.E. Wilson)


Diatom sample preparation followed standard procedures
detailed in Wilson et al. (1994, 1996). Once isolated, diatom
frustules were mounted on glass slides with NaphraxTM
medium. Diatoms were identified and enumerated under oil
immersion using a Leitz DMRB microscope at 1000x magnification with differential interference contrast optics (numerical aperture = 1.30). Over 350 diatom valves were enumerated from each sample. Diatoms were analyzed at 2 cm
intervals, except in the upper 20 cm where analyses were at
1 cm intervals. Taxonomic identifications were based on published descriptions of flora from saline and freshwater lakes
of British Columbia (Cumming et al., 1995), as well as
Krarnmer and Lange-Bertalot (1986, 1988). Past salinity levels were reconstructed from relative abundances of fossil diatoms using a salinity transfer function developed from
219 lakes in southern British Columbia and the Prairies
(Wilson et al., 1996) and the program WACALLB v. 3.3 (Line
et al., 1994). Bootstrapping was used to estimate the standard
error of prediction for each diatom-inferred salinity value.
Diatom-inferred salinities were expressed as the sum of
major ion concentrations in mg.L-l (Wilson et al., 1996), and
are slightly lower than salinities measured as g . ~ -TDS.
l
The
ratio of chrysophycean stomatocysts to diatom frustules was
also used as a qualitative indicator of past salinities (Smol,
1985).

Sedimentary algal pigments (R.D.Vinebrooke)


Sediments for pigment analysis were immediately frozen and
stored in the dark until pigment extraction and analysis.
High-performance liquid chromatography (HPLC) was used
to quantify fossil pigment collcentration~(Leavitt et al., 1989;
Leavitt and Findlay, 1994). Carotenoids, chlorophylls, and
their derivatives were extracted using a standard mixture of

acetone:methanol:water (80: 15:5, by volume) from sediments that had been freeze-dried for 24 h in the dark at a pressure of 0.1 Pa (Leavitt et al., 1989). Pigment concentrations
were quantified using a Hewlett-Packard (HP) Model 1050
HPLC system coupled with a HP Model 1050 photodiode
array detector and a HP Model 1046A scanning fluorescence
detector. Pigments were identified based on spectrophotometric characteristics and by co-chromatography with
authentic standards from the United States Environmental
Protection Agency (Leavitt et al., 1989; Leavitt and Findlay,
1994).
Pigment concentrations were expressed as nmoles
pigment.g-l organic matter (Leavitt, 1993) following determination of organic content by weight-loss-on-ignition at
500' C (Dean, 1974). Recent calibration of annual fossil pigment records against 20 years of phytoplankton data indicates
that organic matter-specific concentrations are significantly
correlated with the abundance of most major algal groups,
and thus permit high-resolution reconstructions of algal
abundance (Leavitt and Findlay, 1994; Leavitt et al., 1994).
Sediments contained more than 50 pigments, many of
which were at low concentration or exhibited infrequent
occurrence. We restricted our analysis to major taxonomically
diagnostic pigments that identified cryptophytes (alloxanthin),
siliceous algae and dinoflagellates (fucoxanthin), diatoms
(diatoxanthin), total cyanobacteria (echinenone), colonial
cyanobacteria (myxoxanthophyll), chlorophytes and
cyanobacteria (lutein-zeaxanthin), and total algal abundance
(p-carotene). Chlorophyll b and pheopigment derivatives
(mainly pheophytin b) were used to distinguish green algae
from cyanobacteria, whose carotenoid, zeaxanthin, was not
separated from the chlorophyte pigment, lutein. Pheophytin
and pheophorbide derivatives of chlorophyll a and chlorophyll b were prepared from pure pigments following the
procedures of Leavitt et al. (1989). The distribution of
carotenoids among algae was detailed in Goodwin (1980).

Phnt macrofossils (R.E. Vance)


Samples were prepared for analysis of plant macrofossils by
gently washing the sediment with tap water through nested
sieves (No. 16,60, and 80 mesh). Macrofossils were removed
from organic residue and identified at 40x magnification
using a stereoscopic microscope. Identifications were made
using Martin and Barkley (1961), Beijerinck (1976), and
Montgomery (1977), and were verified using the Geological
Survey of Canada (GSC) Calgary macrofossil reference collection or vouchered specimens from the University of
Calgary herbarium. Plant macrofossil abundance was
expressed as remains per 100 mL whole sediment.

RESULTS AND DISCUSSION


ChrOnozogyand sedimentation rate
The rate of sediment accumulation was not constant in
Clearwater Lake cores (Fig. 3), indicating that CRS dating
models were most appropriate for estimation of sediment

GSC Bulletin 534

chronology (Appleby and Oldfield, 1983; Turner, 1994; Blais


et al., 1995). Replicate determinations of 2 1 0 ~ activity
b
at
- ~ 28.5
)
cm depth
9.4 cm (0.115 0.004 B ~ . ~ and
(0.053 f0.002 B ~ . ~exhibited
- ~ ) only small errors, suggesting
that estimates of sediment age were reliable (e.g. AD
1983.3 f0.6, 1949.5 f 1.O, respectively). Ages for sediments
deposited before AD 1885 f 5 (below 42.5 cm) were estimated from regressions of sediment depth and 210~b-derived
age, and yielded an approximate basal date of AD 1570 for
core CWS2 at 87 cm. Complete details of the 2 1 0 ~results
b
have been presented by Turner (1994).

Baseline sediment accumulation rates were relatively low


(0.05 g . ~ m - ~ . a -but
l ) increased four-fold during the period
1920-1975, coincident with intensification of European-styleland use (Fig. 3). The most rapid dry mass accumulation occurred during the period 1960-1973. Since 1975, dry
mass accumulation rates have declined, but have remained
two-fold greater than background values. The average linear
sedimentation rate ( S D ) was relatively high (0.65 f 0.27
cm.a-l) but comparable to rates re orted for other northern
Great Plains lakes (0.14-0.95 cm.a- ;Lent et al. (1995), Laird
et al. (1996a), Vinebrooke et al. (1998)).

i'

o
7

Core depth (cm)

2000 1980 1960 1940 1920 1900 1880 1860


Age (years AD)

Figure 3. Unsupported 2 1 0 ~activities


b
in sediments and dry
mass sediment accumulation rates in Clearwater Lake (core
C W S l )as afunction ofA) burial depth and B) estimated age.

Lithostratigraphy
Remarkably few stratigraphic variations were observed in the
major mineralogical components, particle-size attributes, and
organic matter content (Fig. 4). In general, the composition of
late Holocene sediments was similar to that of present-day
offshore deposits (Table 2). Sediments consisted of subequal
proportions of detrital siliciclastic minerals (mainly clay minerals, quartz, plagioclase), detrital carbonate minerals (dolomite, calcite), endogenic and authigenic carbonates (mainly
aragonite), and organic matter. In addition, minor and trace
amounts of pyrite, amphiboles, magnesite, gypsum, and
Na-sulphates occurred sporadically. Sediments were consistently dominated by relatively fine-grained siliciclastics and
detrital carbonates, with some endogenic or authigenic carbonate precipitates. Overall, these patterns provided little
mineralogical or geochemical evidence of greatly altered
salinity during the past approximately 400 years (Fig. 4).
Small stratigraphic changes in the proportion of
magnesian calcite and protodolomite in endogenic and
authigenic mineral fractions suggest that MgICa ratios and
water sources have fluctuated in the recent past, despite comparatively stable total salinity (Fig. 5). For example, the carbonate mineral assemblage of aragonite, protodolomite, and
magnesite below 35 cm indicates high (220) but variable
MgICa ratios and high carbonate alkalinities, with perhaps
low sulphate concentrations, in the lake prior to 1930 (Folk
and Land, 1975; Baker and Kastner, 1981; Last, 1990). In
contrast, magnesian calcite, whose precipitation is encouraged by MgICaratios less than 10 (Eugster and Hardie, 1978),
was restricted to depths of 5-25 cm (ca. 1960-1990). The
occurrence of magnesian calcite only in these sediments indicates that the MgICa ratio of Clearwater Lake has been variable but lower during the past three decades than earlier in the
lake's history. Variation in MgICa ratios most often arises
from changes in relative proportion of water sources (e.g. surface vs. aquifer; Miiller et al. (1972)) and signals alteration in
the hydrological budget of Clearwater Lake during the period
ca. 1960-1 990. Finally, the increased abundance of
protodolomite since 1990 indicates a recent return to present-day MgICa ratios of approximately 20 (Fig. 4, Table 2).
Overall, sediment mineralogy indicates a shift from higher
MgICa ratios before 1930 (220), to lower ratios after 1960
(<lo), and a recent return to pre-1930s values (Fig. 5), each
reflecting changes in water inputs to the lake. Comparison of
these results with an analysis of the previous 3500 years suggests that the recent MgICa composition of Clearwater Lake
has been unusually variable (W.M. Last, unpub. data, 1994).
The appearance of low levels of gypsum at depths of
45-60 cm may indicate a period of increased erosion or sediment redistribution prior to the arrival of ancestral Europeans
(Fig. 5, Evaporites). Gypsum can precipitate in brines rich in
l
Ca-sulphate, but only at salinities greater (i.e. >10 g . ~ -TDS;
Last (1994)) than those in Clearwater Lake at present
(approx. 1 g.L-I TDS) or since ca. 1600 (0.3-0.6 g . ~ - lsee
;
diatoms below). However, because gypsum is common in the
glacial drift of The Missouri Coteau (Freeze, 1969; Wallick
and Krouse, 1977), we infer that sedimentary deposits

P.R. Leavitt et al.

Figure 4. Stratigraphic variation in general physical properties and mineralogy of Clearwater Luke
sediments (core CWS2). 2 1 0 ~chronology
b
is shown on right side of graphs.

Figure 5. Stratigraphic variation in carbonate mineralogy of Clearwater Lake sediments (core


CWS2). 2 1 0 ~chronology
b
as in Figure 4.

GSC Bulletin 534


originated from catchment tills or from reworked shallow-water deposits, and not from endogenic precipitation
during periods of high salinity.
The uniformity of the late Holocene lithostratigraphy in
Clearwater Lake contrasts with the dramatic mineralogical
and geochemical fluctuations that have occurred in other
closed-basin lakes in the region (Last and Schweyen, 1985;
Vance et al., 1992; van Stempvoort et al., 1993; Sack and
Last, 1994; Vance et al., 1997). We attribute the stability of
endogenic and authigenic mineralogy in Clearwater Lake to
its hydrological setting. Because of an absence of surface-water flow at present (Fig. 2), we assume that the hydrological budget of the lake is dominated by groundwater
discharge and recharge, factors that are known to control the
major ion composition, salinity, and chemical sedimentology
processes in other closed-basin lakes (Vance and Last, 1994).
Such a groundwater influence might also moderate the effects
of short-term variations in evaporation and precipitation that
would otherwise regulate lake-water chemistry.

Diatoms
Diatom communities were composed predominantly of benthic taxa, reflecting the shallow nature of Clearwater Lake
during the past 400 years (Fig. 6). Additionally, sedimentary
assemblages consisted almost entirely of diatoms that are
common in fresh (<0.5 ~ . L - ~ T D sto
) subsaline
(0.5-3 g . ~ - l ~ waters
~ S ) (Wilson et al., 1994, 1996;
Cumming et al., 1995). Consequently, overall diatom composition inferred that Clearwater Lake fluctuated within a narrow range of salinity despite periods of drought and
documented water-level change (Fig. 2; Maybank et al.,
1995).
Diatom analyses identified three periods of distinctly different species assemblages. Between ca. 1600 and 1920
(zone I), diatoms consisted mainly of benthic taxa such as
Epithemia argus (Ehrenb.) Kiitz., Amphora cf, pediculus
(Kiitz.) Grun. ex A. Schmidt, Gomphonema angustum
Agardh, Cocconeisplacentula var. euglypta (Ehrenb.) Grun.,
Cocconeis placentula var. lineata (Ehrenb.) V.H.,
Achnanthes minutissima Kiitz, Fragilaria brevistriata Grun.

Figure 6. Stratigraphic variation in diatom taxa in Clearwater Lake sediments (core CWS2), showing taxa
with greater than 5% relative abundance in at least two samples. Salinity optima (g (major ions).~-])for
each taxonfrom Wilson et al. (1996).Tam with the PISCES designation were documentedfrom lakes sampled as part of the Paleolimnological Investigations of Salinity, Climatic, and Environmental Shifts
(PISCES) project (Wilson et al., 1994; Cumming et al., 1995). Diatom zones determined by subjective
analysis.

P.R. Leavitt et al.

in V.H., and Amphora libyca Ehrenb. (Fig. 6, zone 1). The


only common planktonic species, Cyclotella michiganiana
Skvort., also reached its maximum abundance during zone 1,
suggesting that the water column was deep enough to allow
for a simple suspended community. Similarly, chrysophyte
cyst abundance was high relative to that of diatom frustules
(Fig. 6). Chrysophyte assemblages in temperate freshwater
lakes consist primarily of planktonic species (Smol, 1995).
Consequently, high cyst/frustule ratios, combined with some
development of planktonic diatom populations, suggest that
water levels were relatively high.
Striking changes in diatom community composition
occurred shortly after 1920 and persisted until ca. 1977
(Fig. 6, zone 2). The abundances of most common taxa from
zone 1 were reduced ca. 1922 as other species increased for
the first time in this record, including Denticula kuetzingii
Grun., Cyrnbella pusilla Grun. ex A. Schmidt, Cyrnbella cf.
microcephala Grun. in V.H., Cymbella cf, angustata
(W.Sm.) Cleve, and Nitzschia aff, heufleuriana. While some
of these changes suggest a rapid shift towards shallow saline
conditions, the overall pattern is not consistent with major
changes in salinity or climatic conditions. For example, both
the appearance of hyposaline Cymbella pusilla (optimum =
5.5 g . ~ -TDS;
'
Wilson et al. (1996)) and a pronounced drop in
the cyst:diatom ratio (Cumming et al., 1993) suggest that
salinities were elevated during zone 2. Similarly, the loss of
planktonic Cyclotella michiganiana and the decline in the
cyst:diatom ratio possibly resulted from loss of pelagic habitats as the water level declined. However, the majority of taxa
have salinity optima similar to those of species from zone 1
(Wilson et al., 1996), suggesting that community shifts did
not arise from changes in lake-water salinity. Instead, we
infer that these changes in species composition reflect alterations in ionic composition (Fritz et al., 1993; Cumining and
Smol, 1993), such as that inferred from the mineralogical
analysis of Clearwater Lake sediments (Fig. 5). Further,
because diatom community change coincided with increased
activities of ancestral European colonists on the Prairies (Hall
et al., 1999), we infer that human land-use practices are the
main agents of environmental change.
The third distinct diatom community occurred ca. 1977 to
the present (Fig. 6). This period was marked by declines in
Denticula kuetzingii, Cymbella cf. rnicrocephala, Cymbella
cf. angustata, and Cyrnbellapusilla, and increases in the relative abundances of Fragilaria brevistriata, Navicula oblonga
(Kiitz.) Kiitz., Amphora libyca, Cymbella cistula (Ehrenb. in
Hemp. and Ehrenb.) Kirchn., Cymbella sp. 2 PISCES,
Pinnularia rnicrostauron (Ehrenb.) Cleve, and Nitzschia sp. 3
PISCES. Although this assemblage has continued to the present, Navicula oblonga and Fragilaria brevistriata percentages have declined since the mid-1980s. These taxonomic
changes coincided with multi-year drought (1987-1989;
Jones (1991), Wheaton et al. (1992)) and lake-level decline
(Fig. 2).
The diatom data suggest that Clearwater Lake has been
relatively unresponsive to regional droughts during the past
150 years (Fig. 7). For example, severe droughts in western
Canada during 1860 (Bark, 1978; Stockton and Meko, 1983;
Case and MacDonald, 1995) and the early 1890s (Borchert,

1971; Bark, 1978; Jones, 1987) were not recorded by


diatom-inferred salinity. Even the well documented droughts
of the 1930s and 1980s (Evans and Prepas, 1996) resulted in
few changes in diatom-inferred salinity. The observation that
lake levels have recently declined (Fig. 2), yet diatom-inferred salinity has remained constant (Fig. 7) and similar to present-day values (approx. 1 g . ~ -TDS),
'
indicates that
Clearwater Lake is not as sensitive to climatic fluctuations as
are other prairie lakes (e.g. Fritz, 1990;Laird et al., 1996a, b).
Overall, quantitative reconstruction of salinity from diatoms
suggests that concentrations of major ions have been comparatively stable (0.3-0.6 g.L-l TDS), with only slightly elevated
values during 1920-1940,1700-1760, and 1580-1630.
Diatoms may have responded more strongly to changes in
land-use practices and groundwater hydrology than to climate. The most dramatic shifts in diatom community composition coincided with inferred shifts in Mg/Ca ratios, elevated
sediment accumulation rates, and the major influx of European settlers during the period 1906-1921 (Jones, 1987; Hall
et al., 1999). Similarly, the shift in assemblages after ca. 1977
may reflect the end of the main phase of cottage development,

0.05

0.1

0.2

0.5

Diatom-inferred salinity (g c')


Figure 7. Diatom-inferred salinity for the period ca.
AD 1600-present at Clearwater Lake (core CWS2). Salinity
inferences were derived from a salinity transfer function
developed from lakes in southern British Columbia and the
northern Great Plains (Wilson et al., 1996).Inferred salinity
values are mean bootstrap estimates and the errors (dotted
lines) are estimated standard errors of prediction.

GSC Bulletin 534

lake-level decline, or increased landscape development at the


southeast park boundary (Fig. 2). Disturbance of terrestrial
vegetation is known to increase erosional inputs of silicates
and nutrients (Likens et al., 1970), and may have fertilized
benthic diatom assemblages (Fritz, 1989; Hikansson and
RegnC11, 1993; Anderson et al., 1995).

Algal pigments
Concentrations of ubiquitous pigments (e.g. p-carotene) in
Clearwater Lake sediments have been variable during the last
approximately400 years (Fig. 8). The pigment p-carotene is a
reliable marker of total phototroph abundance (Leavitt, 1993;
Leavitt and Findlay, 1994) and its analysis suggested few
directional changes in algal abundance. Instead, striking
shifts in algal composition occurred since ca. 1920, based on
changes in fossil concentrations of taxonomically diagnostic
carotenoids and chlorophylls (Fig. 8). Specifically,
pigments from chrysophytes, diatoms, and dinoflagellates
(fucoxanthin, diatoxanthin) increased five-fold during the
main period of European colonization ca. 1920 (Jones, 1987;
Potyondi, 1995), whereas concentrations doubled for pigments from chlorophytes (chlorophyll b, lutein-zeaxanthin),
cyanobacteria (echinenone, myxoxanthophyll, luteinzeaxanthin), and cryptophytes (alloxanthin). These changes
were likely unrelated to the modest inferred shifts in salinity
l
Fig. 7). Instead, pigment-inferred
(approx. 0.2 g . ~ - TDS;
algal production may have been stimulated by transient
increases in nutrients associated with initial changes in
land-use.
Inferred eutrophication of Clearwater Lake was not sustained, and abundance of total algae, cyanobacteria,
chlorophytes, and cryptophytes apparently declined to

pre-European settlement levels by 1970 (Fig. 8). Similar transient responses of algae to terrestrial disturbances have been
recorded following deforestation (Smol et al., 1983; Sanger,
1988), fires (Schindler et al., 1990), and agriculture (Hurley
et al., 1992; Hall et al., 1999). Unexpectedly, the abundance
of diatoms, chrysophytes, and dinoflagellates are inferred to
have remained two-fold greater than predisturbance levels,
based on continued high concentrations of fucoxanthin and
diatoxanthin. Overall, analysis of fossil pigments suggests
that recent lake conditions have continued to stimulate production of benthic diatoms (Fig. 6, 8).
The recent water quality of Clearwater Lake has remained
variable, with elevated concentrations of most pigments following 1970 (Fig. 8). While these data suggest that recreational or climatic changes since 1970 may have adversely
affected water quality, the fossil patterns are also similar to
artifacts produced by rapid postdepositional pigment degradation (Leavitt, 1993). In particular, the observation that most
pigments show similar patterns of increase in uppermost sediments is consistent with rapid postdepositional losses, and
argues for a cautious interpretation of recent stratigraphic
variations in fossil pigments.

Phnt macrofossils
Macrofossils consisted primarily of stonewort reproductive
structures (oogonia) of the genus Chara (Fig. 9). Concentrations of oogonia began to increase at 20 cm (ca. 1970),peaked
at 13 cm (ca. 1980), and then declined to values two-fold
greater than baseline in the most recent sediments
(1985-present). Charophytes are common in the littoral
zones of a wide variety of lakes (Hutchinson, 1975) and are
considered to be sensitive to environmental change, although

nmoles (pigment) . g-1 (organic matter)

Figure 8. Stratigraphic variation in fossil pigment concentrations of Cleanvater Lake sediments


(core CWS2). Dashed lines indicate sediment age estimatedfrom 2 1 0 ~analysis.
b
Indicator pigments
include p-carotene (all algae), alloxanthin (cryptophytes), fucoxanthin (siliceous algae, some
dinoflagellates), diatoxanthin (diatoms),chlorophyll b (green algae), lutein-zeaxanthin (greenalgae,
cyanobacteria), myxoxanthophyll (colonial cyanobacteria), and echinenone (all cyanobacteria).

P.R. Leavitt et al.

factors controlling their distribution are poorly understood


(Garcia, 1994). A narrow band of thick Chara growth was
observed in relatively shallow water (depth <2 m) near the
northern shoreline of Clearwater Lake (R.E. Vance, unpub.
data, 1993), suggesting that their distribution is limited to
comparatively shallow waters. Unfortunately, it is difficult to
interpret whether elevated charophyte remains signal a
decline in lake level that allowed Chara closer to the coring
site, or transient increases in water during the 1970s (Fig. 2)

0 10

20

40

010

Fossils .I
00 mL-I

Figure 9. Stratigraphic variation in plant macrofossils of


Clearwater Lake sediments (core CWS3). Grey shading representsfive-fold exaggeration of concentration scale.

that increased the extent of the shallow water habitat. Further


research is required to refine paleoecological interpretations
based on remains of Chara.
-

CONCLUSIONS
Analysis of multiple proxy data (mineralogy, diatoms, pigments, and plant macrofossils) identified four main periods
within the past 400 years and documented a transition from
stable freshwater conditions (ca. AD 1600-1920) to shallow
conditions of variable ionic composition (Table 3). The most
recent period (1970-present) was characterized by rapid variations between magnesian calcite (1970-1990) and protodolomite (1990-present; Fig. 5), in lake level (Fig. 2), and in
diatom composition (Fig. 6), although both diatoms and overall mineralogy infer little change in total salinity (Fig. 4,7).
During this period, the abundances of Chara (Fig. 9), siliceous algae (as fucoxanthin; Fig. 8), and chlorophytes (chlorophyll b) were comparatively great, particularly during the
relatively high-water stands of the mid-1970s (Fig. 2). In contrast, the period 1940-1970 was marked by constant intermediate lake levels, relatively low inferred Mg/Ca ratios (<lo),
and declining concentrations of pigments from algae other
than siliceous taxa or dinoflagellates (diatoxanthin,
fucoxanthin; Fig. 8). These conditions may represent a slight
increase in water level and declines in pigments when compared to the period 1920-1940. During that latter era, rural
populations in Saskatchewan reached an historical maximum
(Jones, 1987; Hall et al., 1999). Coincident with these
changes, diatom abundance and species composition were
fundamentally altered (Fig. 6) and pigment-inferred water
quality declined (Fig. 8). Overall, the post-1920 fossil record
suggests a distinctly different history than during the period

Table 3. Summary of the paleolimnological evidence of changing lake conditions in Clearwater Lake, Saskatchewan,
ca. AD 1600-1 993.

Time

Lake
conditions

Lithology,
mineralogy
-

Diatoms
-

Algal pigments

Plant
macrofosslls

- -

1970s- present

Variable lake
level (high in
1970s, low in
1980s)

Mg-calcite
(1970-1 990),
protodolomite
(1990); variable
Mg:Ca ratio

Variable benthic diatom


assemblages

High inferred biomass


(chlorophytes,
siliceous algae)

Increased Chara
abundance
(1970-1 985)

1940-1 970

Intermediate
lake level

Little
protodolomiteor
Mg-calcite; Mg:Ca
ratio e l 0

No planktonic diatoms

Declining algal
biomass; high
diatom abundance

Few

1920-1 940

Low lake
level (late
1930s)

Protodolomite;
Mg:Ca ratio >20

No planktonic species;
shifts in benthic taxa with
land use

Enhancedabundance
of siliceous algae,
chlorophytes,
cyanobacteria

Few

Stable
freshwater
conditions;
high water

Protodolomite;
Mg:Ca ratio >20

Freshwater taxa;
planktonic Cyclotella

Variable algal
abundance; low
diatoms,
chrysophytes, greens

Few

GSC Bulletin 534

1600-1920, during which water levels and MgICa ratios


appeared to have been greater, but diatom and chlorophyte
abundance was reduced (Table 3).
Although closed-basin prairie lakes are regarded as sensitive indicators of climatic change (Fritz, 1996), our findings
suggest that basin-specific factors such as hydrological setting and human land-use history can moderate the effects of
climate on lake chemistry and communities. For example,
Clearwater Lake is uncharacteristically fresh (approx. 1 g . ~ - l
TDS) given the high precipitation deficit (>400 mrn.a-l) and
continued climatic warming (Evans and Prepas, 1996).
Vance and Last (1994) speculated that Clearwater Lake may
be buffered against drought because of groundwater inputs
similar to those reported for Harris Lake, Saskatchewan
(Wilson et al., 1997). In addition, catchment disturbances
arising from modern land-use and recreational activities
appear to have influenced several fossil groups and may have
obscured responses to recent climatic change. Consequently,
we hypothesize that natural environmental perturbations
which influence terrestrial characteristics (e.g. fire) may also
alter lake conditions independent of direct climatic effects.
Fortunately, the use of multiple paleolimnological indicators
can help safeguard against erroneous paleoclimatic interpretations by distinguishing between climatic events and human
activities as controls on lake ecosystem characteristics.

ACKNOWLEDGMENTS
This project was funded by the Geological Survey of Canada
as part of the Palliser Triangle Global Change Project. Additional funding was provided by the Natural Sciences and
Engineering Research Council of Canada (NSERC), as
scholarships to R.D.V. and S.E.W., and as research grants to
P.R.L., W.M.L., and J.P.S. Additional support to R.I.H. was
provided by the University of Regina. Pat Athanasopoulos,
Grant Caresle, Anne Ross, and Chris Teichreb provided valuable assistance in the laboratory.

REFERENCES
Adams, R.H., Rosenwieg, C., Peart, R.M., Ritchie, J.T., McCarl, B.A.,
Glyer, J.D., Curry, R.B., Jones, J.W., Boote, K.J., and Allen, L.H.
1990: Global climate change and US agriculture; Nature, v. 345,
p. 219-221.
Aharonson, E.F., Karasikov, N., Roiterg, M., and Shamir, J.
1986: GALAI-CIS-1 - a novel approach to aerosol particle size analysis;
Journal of Aerosol Science, v. 17, p. 530-536.
Allen, T.
1981: Particle Size Measurement; Chapman and Hall, Toronto, Ontario,
486 p. (third edition).
Anderson, N.J., Renberg, I., and Segerstrom, U.
1995: Diatom production responses to the development of early agriculture in a boreal forest lake-catchment (KassjBn,northern Sweden);
Journal of Ecology, v. 83, p. 809-822.
Appleby, P.G. and Oldfield, F.
1978: Thecalculation of lead-210 dates assuming aconstantrateof supply
of unsupported 210Pbto the sediment; Catena, v. 5, p. 1-8.
1983: The assessment of 210Pb data from sites with varying sediment
accumulation rates; Hydrobiologia,v. 103, p. 29-35.
Baker, P.A. and Kastner, M.
1981: Constraints on the formation of sedimentary dolomite; Science,
V. 213, p. 214-216.

Bark, L.D.
1978: History of American droughts; in North American Droughts, (ed.)
N.J. Rosenberg;American Association for the Advancement of Science, Selected Symposium 15, Westview Press Inc., Boulder, Colorado, p. 9-23.
Beijerinck, W.
1976: Zadenatlas dcr Nederlandsche Flora: Ten Behoeve van de Botanie,
Palaeontologie, Bodemcultuur en Warenkennis; Backhuys and
Meesters, Amsterdam, The Netherlands, 316 p.
Blais, J.M., Kalff, J., Cornett, R.J., and Evans, R.D.
1995: Evaluation of 210Pb dating in lake sediments using stable Pb,
Ambrosia pollen, and 137Cs; Journal of Paleolimnology, v. 13,
p. 169-178.
Borchert, J.R.
1971: The dust bowl in the 1970s; Annals of the Association of American
Geographers, v. 61, p. 1-22.
Canadian National Committee for the International Hydrologic
Decade
1978: Hydrologic Atlas of Canada; Fisheries and Environment Canada,
Ottawa, Ontario, 75 p.
Case, R.A. and MacDonald, G.M.
1995: A dendroclimatic reconstruction of annual precipitation on the
western Canadian Prairies since A.D. 1505 from Pinus flexilis
James; Quaternary Research, v. 44, p. 267-275.
Cumming, B.F. and Smol, J.P.
1993: Development of diatom-based salinity models for paleoclimatic
research from lakes in British ~olumbia(Canada); ~ i d r o b i o l o ~ i a ,
v. 2691270, p. 179-196.
Cumming, B.F., iso on, S.E., HaU, R.I., and Smol, J.P.
1995: Diatoms from British Columbia (Canada) and their relationship to
salinity, nutrients, and other limnological variables; Bibliotheca
Diatomologica Band 31; Gebriider Borntraeger, BerlinIStuttgart,
Germany, 207 p.
Cumming, B.F., Wilson, S.E., and Smol, J.P.
1993: Paleolimnologicalpotential of chrysophyle cysts and scales and of
sponge spicules as indicators of lake salinity; International Journal
of Salt Lake Research, v. 2, p. 87-92.
Dean, W.E.
1974: Determination of carbonate and organic matter in calcareous sediments and sedimentary rocks by loss on ignition: comparison with
other methods; Journal of Sedimentary Petrology, v. 44,
p. 242-248.
El-Ashry, M. and Gibbons, D.C.
1988: Water and Arid Lands of the Western United States; Cambridge
University Press, Cambridge, Massachusetts, 415 p.
Eugster, H.P. and Hardie, L.A.
1978: Saline lakes; in Lakes - Chemistry, Geology, Physics, (ed.)
A. Ler~nan;Springer-Verlag, New York, p. 237-293.
Evans, J.C. and Prepas, E.E.
1996: Potential effects of climate change on ion chemistry and
phytoplankton communities in prairie saline lakes; Limnology and
Oceanography, v. 41, p. 1063-1076.
Folk, R.L. and Land, L.S.
1975: MgICa ratio and salinity: two controls over crystallization of dolomite; American Association of Petroleum Geologists Bulletin,
v. 59, p. 60-68.
Freeze, R.A.
1969: Regional groundwater flow - Old Wives Lake drainage basin,
Saskatchewan; Department of Energy, Mines and Resources,
Canada, Inland Waters Branch, Scientific Series, v. 5,245 p.
Fritz, S.C.
1989: Lake development and limnologicalresponse to prehistoric and historic land-use in Diss, Norfolk, U.K.; Journal of Ecology, v. 77,
p. 182-202.
1990: Twentieth-century salinity and water-level fluctuations in Devils
Lake, North Dakota: test of a diatom-based transfer function; Limnology and Oceanography, v. 35, p. 1771-1781.
1996: Paleolimnological records of climatic change in North America;
Limnology and Oceanography, v. 41, p. 882-889.
Fritz, S.C., Engstrom, D.R., and Haskell, B.J.
1994: 'Little Ice Age' aridity in the North American Great Plains: a high
resolution reconstruction of salinity fluctuations from Devils Lake,
North Dakota, USA; The Holocene, v. 4, p. 69-73.

P.R. Leavitt et al.

Fritz, S.C., Juggins, S., and Battarbee, R.W.


1993: Diatom assemblages and ionic characterization of lakes of the
northern Great Pla~ns,North America: a tool for reconstructing past
salinity and climate fluctuations;Canadian Journal of Fisheries and
Aquatic Sciences, v. 50, p. 1844-1856.
Fritz, S.C., Juggins, S., Battarbee, R.W., and Engstrom, D.R.
1991: Reconstruction of past changes in salinity and climate using a diatom-based transfer function; Nature, v. 352, p. 706-708.
Garcia, A.
1994: Charophyta: their use in paleolimnology; Journal of
Paleolimnology,v. 10, p. 43-52.
Gasse, F., Juggins, S., and Ben Khelifa, L.
1995: Diatom-based transfer functions for inferring past hydrochemical
characteristics of African lakes; Palaeogeography,
Palaeoclimatology,Palaeoecology, v. 117, p. 31-57.
Gleick, P.H.
1989: Climate change, hydrology, and water resources; Reviews in Geophysics, v. 27, p. 329-344.
Goldsmith, J.R. and Graf, D.L.
1958: Relation between lattice constants and composition of the Ca-Mg
carbonates; American Mineralogist, v. 43, p. 84-101.
Goodwin, T.W. (ed).
1980: The Biochemistry of the Carotenoids, Volume 1. Plants; Chapman
and Hall, London, United Kingdom, 377 p.
Gullett, D.W. and Skinner, W.R.
1992: The state of Canada's climate: temperature change in Canada
1985-1991; Environment Canada, Atmospheric Environment
Service, State of the Envkonment Report No. 92-2,36 p.
HBkansson, H. and RegnCII, J.
1993: Diatom succession related to land use during the last 600 ycars: a
study of a small eutrophic lake in southern Sweden; Journal of
Paleolimnology, v. 8, p. 49-69.
Hall, R.I., Leavitt, P.R., Quinlan, R., Dixit, A.S., and Smol, J.P.
1999: Effects of agriculture, urbanization and climate on water quality in
the northern Great Plains; Limnology and Oceanography, v. 43,
p. 739-756.
Hammer, U.T.
1990: The effects of climate change on the salinity, water levels and biota
of Canadian prairie saline lakes; Verhandlungen der Internationale
Vereinigung fiir Angewandte und Theoretische Limnologie, v. 24,
p. 321-326.
Hammer, U.T., Shamess, J., and Haynes, R.C.
1983: The distribution and abundance of algae in saline lakes of
Saskatchewan;Hydrobiologia, v. 105, p. 1-26.
Hardy, R. and Tucker, M.
1988: X-ray diffraction of sediments; it1 Tcchniques in Sedimentology,
(ed.) M. Tucker; Blackwell Scientific Publishers, Boston, Massachusetts, p. 191-228.
Hurley, J.P., Armstrong, D.E., and DuVall, A.L.
1992: Historical interpretation of pigment stratigraphy in Lake Mendota
sediments; in Food Web Management, (ed.) J.F. Kitchell;
Springer-Verlag,New York, p. 49-68.
Hutchinson, G.E.
1975: A Treatise on Limnology, Volume 111, Limnological Botany;
John Wiley and Sons, New York, 660 p.
Jones, D.C.
1987: Empire of Dust; University of Alberta Press, Edmonton, Alberta,
316 p.
Jones, K.H.
1991: Drought on the Prairies; in Symposium on the Impacts of Climatic
Change and Variability on the Great Plains (1990: Calgary,
Alberta), (ed.) G. Wall; Department of Geography, University of
Waterloo, Waterloo, Ontario, p. 125-129.
Klug, H.P. and Alexander, L.E.
1974: X-ray Diffraction Procedures for Polycrystalline and Amorphous
Materials; John Wiley and Sons, New York, 346 p.
Krammer, K. and Lange-Bertalot, H.
1986: Bacillariophyceae. I. Teil: Naviculaceae; in Siiflwasserflora von
Mitteleuropa, Band 2/l, (ed.) H. Ettl, J. Gerloff, H. Heynig, and
D. Mollenhauer;GustavFischerVerlag, Stuttgart,Germany, 876 p.
1988: Bacillariophyceae. 2. Teil: Bacillariaceae, Epithemiaceae,
Surirellaceae; in SiiBwasserflora von Mittelcuropa, Band 212, (ed.)
H. Ettl, J. Gerloff, H. Heynig, and D. Mollenhauer; Gustav Fischer
Verlag, Stuttgart, Germany, 61 1 p.

Laird, K.R., Fritz, S.C., Grimm, E.C., and Mueller, P.O.


1996a: Century-scale paleoclimatic reconstruction from Moon Lake, a
closed-basin lake in the northern Great Plains; Limnology and
Oceanography,v. 41, p. 890-902.
Laird, K.R., Fritz, S.C., Maasch, K.A., and Cumming, B.F.
1996b: Greater drought intensity before AD 1200 in the northern Great
Plains, USA; Nature, v. 384, p. 552-554.
Last, W.M.
1990: Lacustrine dolomite - an overview of modern, Holocene, and Pleistocene occurrences; Earth Science Reviews, v. 27, p. 221-263.
1994: Paleohydrology of playas in the northern Great Plains: perspectives
from Palliser's Triangle; in Paleoclimate and Basin Evolution of
Playa Systems, (ed.) M.R. Rosen; Geological Society of America,
Special Paper 289, p. 69-80.
Last, W.M. and Schweyen, T.H.
1985: Late Holocene history of Waldsea Lake, Saskatchewan, Canada;
Quaternary Research, v. 24, p. 219-234.
Leavitt, P.R.
1993: A review of factors that regulate carotenoid and chlorophyll deposition and fossil pigment abundance; Journal of Paleolimnology,v. 9,
p. 109-127.
Leavitt, P.R. and Findlay, D.L.
1994: Comparison of fossil pigments with 20 years of phytoplankton data
from eukophic Lake 227, Experimental Lakes Area, Ontario;
Canadian Journal of Fisheries and Aquatic Sciences, v. 51,
p. 2286-2299.
Leavitt, P.R., Carpenter, S.R., and Kitchell, J.F.
1989: Whole-lake experiments: the annual record of fossil pigments and
zooplankton; Limnology and Oceanography,v. 34, p. 700-717.
Leavitt, P.R., Hann, B.J., Smol, J.P., Zeeb, B.A., Christie, C.E.,
Wolfe, B., and Kling, H.J.
1994: Paleolimnological analysis of wholelake experiments: an overview of results from Experimental Lakes Area Lake 227; Canadian
Journal of Fisheries and Aquatic Sciences, v. 51, p. 2322-2332.
Lemmen, D.S., Dyke, L.D., and Edlund, S.A.
1993: The Geological Survey of Canada's Integrated Research and Monitoring Area (IRMA) projects: a contribution to Canadian global
change research; Journal of Paleolimnology, v. 9, p. 77-83.
Lent, R., Lyons, W.B., Showers, W.J., and Johannesson, K.H.
1995: Late Holocene paleocli~naticand paleobiologic records from sediments of Devil's Lake, North Dakota; Journal of Paleolimnology,
v. 13, p. 193-207.
Likens, G.E., Bormann, F.H., Johnson, N.M., Fisher, D.W., and
Pierce, R.S.
1970: Effects of forest cutting and herbicide treatment on nutrient budgets
in Hubbard Brook watershed-ecosystem;Ecological Monographs,
v. 40, p. 2347.
Line, J.M., ter Braak, C.J.F., and Birbs, H.J.B.
1994: WACALIB version 3.3 - a computer program to reconstruct environmental variables from fossil assemblagesby weighted averaging
and to derive sample-specific errors of prediction; Journal of
Paleolimnology, v. 10, p. 147-152.
Marino, R., Howarth, R.W., Shaness, J., and Prepas, E.E.
1990: Molybdenum and sulfate as controls of the abundance of nitrogen-fixing cyanobacteria in saline lakes in Alberta; Limnology and
Oceanography, v. 35, p. U5-259.
Marquart, R.G.
1986: PDSM: mainframe searchlmatch on an IBM PC; Powder Diffraction, v. 1, p. 34-36.
Martin, A.C. and Barkley, W.D.
1961: Seed Identification Manual; University of California Press,
Berkeley, California, 220 p.
Maybank, J., Bonsal, B., Jones, K., Lawford, R., O'Brien, E.G.,
Ripley, E.A., and Wheaton, E.
1995: Drought as a natural disaster; Atmospheric-Ocean, v. 33,
p. 195-222.
Meko, D.M.
1992: Dendroclimaticevidence from the Great Plains of the United States;
in Climate Since AD 1500, (ed.) R.S. Bradley and P.D.Jones;
Routledge, New York, p. 312-330.
Montgomery, F.H.
1977: Seeds and Fruits of Plants of Eastern Canada and Northeastern
United States; University of Toronto Press, Toronto, Ontario, 232 p.
Miiller, G., Irion, G., and Forstner, U.
1972: Formation and diagenesis of inorganic Ca-Mg carbonates in the lacustrine environment; Natiirwissenschaften,v. 59, p. 159-164.

GSC Bulletin 534


Potyondi, B.
1995: In Palliser's Triangle: Living in the Grasslands 1850-1930; Purich
Publishing, Saskatoon, Saskatchewan, 143 p.
Sack, L.A. and Last, W.M.
1994: Lithostratigraphy and recent sedimentation history of Little Manitou Lake, Saskatchewan,Canada; Journal of Paleolimnology,v. 10,
p. 199-212.
Sanger, J.E.
1988: Fossil pigments in paleoecology and paleolimnology; Palaeogeography, Palaeoclimatology,Palaeoecology, v. 62, p. 343-359.
Schindler, D.W., Beatty, K.G., Fee, E.J., Cruikshank, D.R.,
DeBruyn, E.R., Findlay, D.L., Lindsay, G.A., Shearer, J.A.,
Stainton, M.P., and Turner, M.A.
1990: Effects of climatic warming on lakes of the central boreal forest;
Science, v. 250, p. 967-970.
Schnltz, L.G.
1964: Quantitative interpretation of mineralogical composition from
X-ray and chemical data for thePierre Shale; United States Geological Survey, Professional Paper 391-C, 31 p.
Smol, J.P.
1985: The ratio of diatom frustules to chrysophycean statospores: a useful
paleolimnological index; Hydrobiologia, v. 123, p. 199-208.
1995: Application of chrysophytes to problems in paleoecology; in
Chrysophyte Algae: Ecology, Phylogeny and Development, (ed.)
C.D. Sandgren, J.P. Smol, and J. Kristiansen; Cambridge University Press, Cambridge, United Kingdom, p. 303-330.
Smol, J.P., Brown, S.R., and McNeely, R.N.
1983: Cultural disturbancesand trophic history of a small mero~nicticlake
from central Canada; Hydrobiologia, v. 103, p. 125-130.
Stockton, C.W. and Meko, D.W.
1983: Drought recurrence in the Great Plains as reconst~uctedfrom
long-term tree-ring rccords; Journal of Climate and Applied Meteorology, v. 22, p. 17-29.
Teller, J.T. and Last, W.M.
1990: Paleohydrologicalindicators in playas and salt lakes, with examples
from Canada, Australia, and Africa; Palaeogeography,
Palaeoclimatology,Palaeoecology, v. 76, p. 215-240.
Torgersen, T., DeDekker, P., Chivas, A.R., and Bowler, J.M.
1986: S a l t lakes: a discussion of processes influencing
paleoenvironmental interpretation and recommendations; Palaeogeography, Palaeoclimatology,PaIaeoecology, v. 54, p. 7-19.
Turner, L.J.
1994: 2'0Pb dating of lacustrine sediments from Clearwater Lake (Core
064, Station CWSl), Saskatchewan;National Water Research Institute, Burlington, Ontario, Contribution 94-141.27 p.
Vance, R.E.
1997: The Geological Survey of Canada's Palliser Triangle Global
Change Project: a multidisciplinary geolimnological approach to
predicting potential global change impacts on the northern Great
Plains; Journal of Paleolimnology,v. 17, p. 3-8.
Vance, R.E. and Last, W.M.
1994: Paleolimnology and global change on the southern Canadian
Prairies; in Current Research 1994-B; Geological Survey of
Canada, p. 49-58.

Vance, R.E. and Last, W.M. (cont.)


1996: Stop 31: Clearwater Lake; in Landscapes of the Palliser Triangle,
Guidebook for the Canadian Geomorphology Research Group Field
Trip, (ed.) D.S. Lemmen; Canadian Association of Geographers,
1996 Annual Meeting, Saskatoon, Saskatchewan,p. 78-79.
Vance, R.E. and Mathewes, R.W.
1994: Deposition of modern pollen and plant macroremains in a
hypersaline prairie lake basin; Canadian Journal of Botany, v. 72,
p. 539-548.
Vance, R.E., Last, W.M., and Smith, A.J.
1997: Hydrologicand climatic implicationsof a multidisciplinarystudy of
Holocene sediment from Kenosee Lake, southeastern
Saskatchewan, Canada; Journal of Paleolimology, v. 18,
p. 365-393.
Vance, R.E., Mathewes, R.W., and Clague, J.J.
1992: 7000 year record of lake-level change on the northern Great Plains:
a high-resolution proxy of past climate; Geology, v. 20, p. 879-882.
van Stempvoort, D.R., Edwards, T.W.D., Evans, M.S.,
and Last, W.M.
1993: Paleohydrology and paleoclimate records in a saline prairie lake
core: mineral, isotope and organic indicators; Journal of
Paleolimnology, v. 8, p. 135-147.
Vinebrooke, R.D., Hall, R.X., Leavitt, P.R., and Cumming, B.F.
1998: Fossil pigments as indicators of phototrophic response to salinity
and climatic change in lakes of western Canada; Canadian Journal
of Fisheries and Aquatic Sciences, v. 55, p. 668-681.
Wallick, E.I. and Krouse, H.R.
1977: Sulfur isotope geochemistry of a groundwater-generated
Na2S04/Na2C03 deposit and the associated drainage basin of
Horseshoe Lake, Metiskow, east-central Alberta, Canada; in The
Second International Symposium on Water-Rock Interaction,
Strasbourg, France, p. 1156-1 164.
Wheaton, E.E., Arthur, L.M., Chorney, B., Shewchuk, S., Thorpe, J,
Whiting, J., and Wittrock, V.
1992: The Prairie drought of 1988; Climatological Bulletin, v. 26,
p. 188-205.
Wiche, G.J.
1986: Hydrologic and climatologic factors affecting water levels in
Devil's Lake, North Dakota; United States Geological Survey,
Water Research Investigations Report 86-4320,86 p.
Will, G.F.
1946: Tree ring studies in North Dakota; North Dakota Agricultural
Experiment Station, Bulletin 338, p. 1-24.
Wilson, S.E., Cumming, B.F., and Smol, J.P.
1994: Diatom-salinity relationships in 111 lakes from the Interior Plateau
of British Columbia, Canada: the development of diatom-based
models for paleosalinity reconstructions; Journal of
Paleolimnology, v. 12, p. 197-221.
1996: Assessing the reliability of salinity inference models from diatom
assemblages: an examination of a 219 lake data set from western
North America; Canadian Journal of Fisheries and Aquatic
Sciences, v. 53, p. 1580-1594.
Wilson, S.E., Smol, J.P., and Sauchyn, D.J.
1997: A Holocene paleosalinity diatom record from southwestern
Saskatchewan, Canada: Harris Lake revisited; Journal of
Paleolimnology, v. 17, p. 23-31.

A postglacial plant macrofossil record of vegetation


and climate change in southern Saskatchewan
Catherine H. ~ a n s a and
' James F. ~ a s i n ~ e r ~
Yansa, C.H. and Basinger, J.F., 1999: A postglacial plant macrofossil record of vegetation and
climate change in southern Saskatchewan; in Holocene Climate and Environmental Change in
the Palliser Triangle: A Geoscientific Contextfor Evaluating the Impacts of Climate Change on
the Southern Canadian Prairies, (ed.) D.S. k m m e n and R.E. Vance; Geological Survey of
Canada, Bulletin 534, p. 139-1 72.

Abstract: Plant macrofossil analyses and 14cages from a small, closed-drainage basin on The Missouri
Coteau are used to reconstruct six phases of postglacial environmental change in southern Saskatchewan.
Fossil remains of 41 taxa of vascular and nonvascular plants have been recognized and are illustrated.
Sparse macrofossils recovered from the basal diarnicton (phase 1) are likely redeposited. Abundant fossils indicate establishment of an open, white spruce forest by ca. 10 200 BP (phase 2), followed by development of a pond and replacement of the spruce forest by deciduous parkland vegetation (phase 3) that
persisted until ca. 8800 BP. The pond shallowed at the end of this period. Prairie fires are evident between
ca. 8800 and 7700 BP (phase 4). Water levels rose between ca. 7700 and 5800 BP, and a semipermanent
prairie pond was established (phase 5). After 5800 BP, this wetland became ephemeral (phase 6) and no longer conducive for fossil preservation.

R6sumC : Des

analyses de macrofossiles v6gCtaux et des datations au 14c dans un petit bassin


hydrographique fermC sur le coteau du Missouri ont Ctt utilisees pour reconstituer six phases de
changements environnementaux postglaciaires dans le sud de la Saskatchewan. Les vestiges fossiles de
41 taxons de plantes vasculaires ont kt6 reconnus et sont illustrb.
Les rares macrofossiles prClev6 dans le diamicton de base (phase 1)sont vraisemblablement red6poses.
La presence de fossiles abondants indique qu'une for& claire d'kpinettes blanches s'Ctait implantCe d b
environ 10 200 BP (phase 2), aprks quoi il y a eu dtveloppement d'un Ctang et remplacement de la for&t
d'tpinettes par une vCgCtation de forbt-parc i feuilles caduques (phase 3) qui a persist6 jusqu'i environ
8 800 BP. A la fin de cette periode, 1'Ctang est devenu moins profond. L'action des feux de prairies est
manifeste d'environ 8 800 B 7 700 BP (phase 4). Le niveau de l'eau a monte entre environ 7 700 et 5 800 BP
et un Ctang de prairie semi-permanent est apparu (phase 5). Aprbs 5 800 BP, ces terres hurnides sont
devenues CphCmbres (Phase 6) et dCfavorables B la conservation des fossiles.

Department of Geography, University of Wisconsin-Madison, Madison, Wisconsin, U.S.A. 53706-1491


Department of Geological Sciences, University of Saskatchewan, Saskatoon, Saskatchewan S7N 5E2

GSC Bulletin 534

INTRODUCTION
Meteorological records for the Canadian Prairies are available for only the last centuly (Gullett and Skinner, 1992) and
provide insufficient evidence to distinguish between climate
change and meteorological variability. Therefore, numerous
sites with detailed environmental records spanning the past
several millennia have been investigated within this region by
participants in the Palliser Triangle Global Change Project, in
order to identify long-term trends in regional climate
(Lemmen et al., 1993; Last, 1994; Vreeken, 1994; Vance and
Last, 1994; Yansa, 1995; Beaudoin, 1996; Vance, 1997;
Wilson et al., 1997).
Reconstruction of past vegetation provides valuable
information concerning changes in climate and other environmental factors. Fossil pollen studies have been used to
reconstruct past regional vegetation; however, long-distance
dispersal and poor taxonomic control limit the resolution of
such interpretations (Faegri et al., 1989; Vance et al., 1994).
Alternatively, accurate information on past local vegetation
assemblages and environments may be provided by the study
of plant macrofossils, including wood, leaves, seeds, and
fruits, because most macrofossil remains of Holocene age can
be identified to the species level, and the majority of seeds
and other p l a n t remains a r e a u t o c h t h o n o u s and
parautochthonous (Birks and Birks, 1980; Bateman, 1991;
Vance and Mathewes, 1994).
Coring of lake sediments is the most common method of
collecting plant fossil remains; however, suitable lakes are
uncommon on the dry northern Great Plains (Barnosky et al.,
1987; Vance et al., 1993). Nevertheless, the available lake
data do furnish a good regional paleovegetation record for the
last 5000-6000 years (Mott, 1973; Last and Slezak, 1988)
and, in a few cases, for the last 9000-10 000 years (Vance and
Last, 1994; Vance, 1997). A number of recent studies in the
Palliser Triangle region of southern Saskatchewan and southeastern Alberta indicate that significant paleoenvironmental
data can also be obtained from postglacial sediments that fill
small (30-80 m diameter) kettle depressions in hummocky
and ice-thrust moraine (Beaudoin, 1992; Klassen, 1994;
Yansa, 1998). The sediments that accumulated in these kettles often have exquisitely preserved plant and animal
remains that provide a record of floral, climatic, and hydrological changes from immediate postglacial to mid-Holocene
time. The fossil records obtained from kettle-fill sites, therefore, complement the paleoenvironmental interpretations
provided by lake records from this region (e.g. Last and
Slezak, 1986; Sauchyn and Sauchyn, 1991; Vance and Last,
1994; Vance, 1997).
Data on kettle-fill sites were first reported in
Saskatchewan by Dew (1959), Kupsch (1960), DeVries
(1963), Ritchie and DeVries (1964), Ritchie (1966), andMott
and Christiansen (1981); and in North Dakota by Moir
(1957), Thompson (1962), McAndrews et al. (1967),
Cvancara et al. (1971), and Malo (1988). These studies were
largely preliminary with poor chronological control. A

140

number of kettle-fill deposits have been recently investigated by


Klassen (1994) in southwestern Saskatchewan and by Beaudoin
(1992, 1996) in southeastern Alberta, but chronostratigraphlc
resolution of these studies was also limited.
This study provides a history of the local vegetation and
climate change following deglaciation on The Missouri
Coteau upland of southern Saskatchewan, from ca.
10 200-5800 BP, based on plant macrofossils from a kettle
fill. High-resolution sampling, species-level determinations,
and several radiocarbon ages have allowed determination of
the time and duration of the local vegetational, climatic, and
environmental changes, and allow for comparison with other
sites in the region.

SETTING
The Andrews site (lat. 50'20' N, long. 105'52' W; elev.
720 m a.s.l.), located on The Missouri Coteau upland about
100 km southwest of Regina, Saskatchewan (Fig. I), is typical of
the closed-drainage kettles of that region.

Geology
The Missouri Coteau upland (Fig. 1) is a northwest-trending
band of predominantly hummocky moraine, averaging 50 km
in width and extending nearly 1300 km from south-central
South Dakota to west-central Saskatchewan (Acton et al.,
1960; Clayton, 1967). Bedrock, mainly nonmarine deposits
of the Paleocene Ravenscrag Formation, is overlain by up to
100 m of Pleistocene sediments (Christiansen, 1961; Klassen,
1992). The surficial glacial sediments on much of The
Missouri Coteau were deposited during a period of ice
stagnation, forming the characteristic 'knob-and-kettle'
hummocky terrain seen throughout most of this upland (Clayton,
1967). The closed drainage of these kettles makes them ideal
traps for lacustrine, eolian, and colluvial sediments (Hubbard
and Linder, 1986).

Vegetation
The Andrews site is situated within the northern mixed-grass
prairie region (Stipa-Bouteloua association), which in
Canada extends as an arc from southern Alberta through
Saskatchewan to Manitoba (Risser et al., 1981). Eighty percent of the vegetation of this region consists of associations
dominated by Agropyron (wheatgrass), Stipa (needle grass),
and Bouteloua (grama grass) species (Barker and Whitman,
1988). Woody species such as Salix spp. (willow) and
Populus spp. (poplar) occur along riverbanks, coulees, and
moist depressions. Small stands of Populus trernuloides
(aspen poplar) are found near the bottom of a few kettles in
the immediate vicinity of the study site, indicating that
groundwater seepage is locally sufficient to support some
trees on an otherwise treeless landscape.

C.H. Yansa and .l.F. Basinger

Figure 1. Location of the Andrews site ( I ) in the central Palliser Triangle (bold dashed line). Light
shading indicates Brown Chernozemic Soil Zone; darlcshading indicates The Missouri Coteau. Locations ofkettle-fillfossi sites: l )Andrews; 2) Hafichuk; 3 ) Kyle; 4)Beechy; 5)Herbert; 6) Crestwynd;
7)Scrimbit; 8) Wrusu; 9) Val Marie; 10)Horseman; 11) Fletcher; 12)Jenner; and 13) Webb. Radiocarbon ages for the Andrews site are listed in Table 1; locations, 14cages, and references for additional sites are listed in Table 2.

Climate
The present regional climate is strongly continental, with
extreme differences between summer and winter temperatures and low annual precipitation. Evapotranspiration rates
throughout this area are high, with an annual moisture deficit
greater than 300 rnrn (Winter, 1989). It is not surprising,
therefore, that extended periods of dry weather or droughts
are common on the northern Great Plains in Canada. The
closest meteorological station is at Moose Jaw, located 22 krn
east of the Andrews site, which reports an average frost-free
period of 105 days, January mean temperature of -14.2"C,
July mean temperature of +19.7OC, and mean annual precipitation of 357 mm (Environment Canada, 1993).

Hydrology
Each closed kettle within the hummocky moraine is dependent upon snowmelt runoff, which accumulates as standing
water and slowly recharges the underlying aquifer from
spring to mid-summer through the poorly permeable
clay-rich surficial sediments (Lissey, 1971). This groundwater recharge-dominated regime results in kettle water
chemistry being fresh to slightly brackish. Analysis by
Christiansen (1961) of water samples collected about 4 m
below the water table from three test holes near our study site
indicated that total dissolved solids (TDS) ranged from 880 to
1360 mg.L-l. The areal extent of surface waters expands and
contracts seasonally and varies considerably with annual

changes in precipitation and temperature. Today, many kettles on The Missouri Coteau are ephemeral ponds; however,
the plant macrofossil record from the Andrews site suggests
that water levels were higher and perennial in the past.

METHODS
Field sampling
Sediment samples were collected from a 5.8 m section where
well preserved trunks of Picea glauca (white spruce) were
found in situ after excavation for a livestock watering-hole
(Fig. 2). Samples were taken at 5 cm intervals and stored in
labelled plastic bags at 4' C to prevent microbial and fungal
decomposition prior to processing (Warner, 1990; Gale and
Hoare, 1991). Sixty-seven samples from the plant
macrofossil-bearing beds (3.1-5.1 m) and the underlying
diarnicton (5.1-5.8 m) form the basis for this study (Fig. 3)
and are late Pleistocene to mid-Holocene in age. Samples collected from the upper 3.05 m of the section lackedidentifiable
plant macrofossils (Fig. 3).

Sample preparation and analysis


A subsample of 50 cm3 of sediment from each sample was
processed for macrofossils. Each was treated with approximately 900 mL of distilled water and 20 rnL of Quaternary

GSC Bulletin 534

OTMdetergent. The mixture was gently heated to 40C and


agitated for 2 h on an oscillating hot plate to deflocculate the
clay minerals, then washed with water through nested screens
of 1 rnrn, 250 pn, and 125 pn mesh sizes. Following the
methods of A. Beaudoin (pers. comm., 1994),the fossils were
stored in labelled glass vials filled with a 1:1:1 mixture of
95% ethanol, distilled water, and glycerine to which three
a
phenol
were added t' prevent
and microbial decomposition. Samples are stored in the

University of Saskatchewan Paleobotanical Collection


(USPC), Department of Geological Sciences, University of
Saskatchewan, Saskatoon.
Fossil seeds and fruits were identified with the aid of keys
md pubbshhe Ereporn(Mh md BBarkley,1961; Montgomeq,
1977; LeVesqueet al., 1988), Species identifications were
confirmed by
with a
seed and fruit collection prepared for this study and presently housed in the

Figure 2.
Photographs of the Andrews study site:
A) dragline excavation of site for a livestock
waterinn-hole: GSC 1999-044A B ) 5.8 m section sampled along one wall of the dugout at the
Andrews site; field assistants (indicated by
arrow) for scale; GSC 1999-944B C) close-up
of the 3.8 m long trench seen in the lowerpart of
Figure 2B, showing buried plant macrofossilbearing layers; in situ trunks of Picea glauca
overlie roots that penetrate the diamicton;
white arrow points to the contact between the
lacustrine sediments (zone III) and overlying
charcoal-rich sandy clay (zone IV) at 400 cm;
trowel in lower right corner for scale. GSC
1999044C. Photographs by C.H. Yansa.

C.H. Yansa and J.F. Basinger

Fraser Herbarium, University of Saskatchewan. Identifications of nonvascular plants were provided by R. Belland,
Devonian Botanic Gardens, Edmonton, Alberta.
Fossil fruits, seeds, and vegetative organs were mounted
on scanning electron microscope (SEM) stubs with double-sided tape and coated with gold. Specimens were viewed
and photographed on a Philips 505 SEM at 30 kV. Conifer
leaves with cuticle were first soaked in CalgonTM-softened
water (sodium hexametaphosphate solution) for 48 h, oxidized for up to one minute in household bleach (6% sodium
hypochlorite) until clear, and then flushed with
CalgonTM-softenedwater (method of J. Skog, as cited in
LePage and Basinger (1995)). The cuticle was easily separated from mesophyll using insect pins, and examined and
photographed under an SEM.

The relative abundance of leaves, mosses, and charcoal


was estimated using a scale of 0 to 5 , as proposed by Birks and
Birks (1980) and Warner (1990), whereas seeds and fruits
were counted. These abundance numbers were then plotted
(Fig. 4) using TiliaTMand Tilia.graphTM(Grimm, 1992).
Six 14cages for the Andrews site were obtained from
uncharred Picea wood using conventional dating methods,
and from nonaquatic seeds, charcoal flakes, and plant fragments using accelerator mass spectrometry (AMS) techniques (Table 1). Samples were selected from layers where
significant plant macrofossil and stratigraphic changes were
noted (Fig. 3). Two of the wood samples were from trunks
buried in laminated sediments, and the third was from an in
situ Picea root penetrating a litter layer into diamicton
(Fig. 2C; Table 1). The seeds and plant remains selected for
AMS 14C age determinations showed no signs of abrasion.
Preparation of these fossils included rinsing with distilled
water (neither Quaternary OTMnor heating was used in sample processing). The radiocarbon ages (Table 1) have been
corrected for 613c fractionation (R. McNeely, pers. comm.,
1995).

Taxonomy
All fossil remains were compared with extant species from a
comparative collection of 810 prairie, parkland, and forest
taxa. Most of the fossils were identified to species level. For a
few poorly preserved reproductive and vegetative remains,
assignment to a specific taxon is tentative and indicated with
'cf.' in the discussion that follows. The nomenclature and terminology f o r the Lemnaceae, Najadaceae, and
Chenopodiaceae follow the taxonomic revisions of V. Harms
(Saskatchewan Floral Project, unpub. data, 1994) and, for the
remaining taxa, Looman and Best (1987). Unless otherwise
noted, the range and habitat of modern analogues are based on
Johnson (1961), Harms et al. (1986), Looman and Best
(1987), and V. Harms (Saskatchewan Floral Project, unpub.
data, 1994). Scanning election microscope photographs of
the 41 plant macrofossil taxa identified in this study are
included in Appendix A. Descriptions of, and remarks on, the
fossils are given in Appendix B.

LEGEND

@
3.9 m

Poplar leaf

c----o Spruce log

Table 1. Radiocarbon ages (corrected for 613Cfraction-

ation) of significant plant macrofossil-bearing levels of the


Andrews site.

4.0 m
Depth (cm)

Charcoal-rich
sandy clay

310-315

silty clay
5.0m
5.1 m

Dating lab-

Sandy clay

Litter

sample #
TO-5018

"C age (a BP);


dating method

5770 * 80;

Materials dated

Charcoal flakes and plant motlets

AMS
390-395

TO-4780

405-41 0

TO-5019

7670 i- 80;
AMS
8790 * 140;

AECV-2048C

AMS
10 200

445450

* 140;

Charcoal flakes and plant rootlets


Chenopodium salinum seeds and
Scirpus amerlcanusfruits
Picea wood (trunk)

conventional

Diamicton
5.8m

Figure 3. Stratigraphic column of the Andrews site, illustrating the vegetation zones, j4c ages, lithology, and fossils.

490495

GSC-5822

10 200 i 80;

conventional

Picea wood (trunk)

Picea wood (root)


10 230 * 140;
conventional
Age-dating laboratories:TO, IsoTrace Laboratory, University of Toronto; AECV,
Alberta Environmental Centre-Vegreville; GSC, Geological Survey of
500505

AECV-2047C

Canada-Ottawa.

GSC Bulletin 534

RESULTS
The results of this study enable reconstruction of theenvironmental setting and local vegetation, from the time of establishment of vegetation following deglaciation to the
mid-Holocene. Based on changes in plant macrofossil assemblages, five zones were
(Fig. 3, 4). These zones
are associated with I4cages and often correlate with changes
in lithology.

Zone I - reworked fragments; diumicton


(580-51 0 cm)
The base of the section consists of a very dark grey (5Y 311,
MunsellTMColour Chart) diamicton composed of a calcareous clay loam with striated cobbles (Fig. 3). Rare and poorly
preserved plant macrofossils were recovered (Fig. 4),
although these remains are likely reworked. Abrasive degradation of most of the specimens indicates transport, while
those taxa identified as aquatics are not typical of an early
succession vegetational assemblage.

Zone I1 - Picea; litter and laminated silty clay


(510-445 cm)
In situ Picea sp. roots (Fig. 2C), which penetrate a 10 cm thick
litter layer (Fig. 3, 4) containing abundant spruce
macrofossils and the seeds and fruits of shrubby and

herbaceous angiosperms, indicate the establishment of a


spruce forest by ca. 10 200 BP (Table 1). Cone and seed-wing
morphology were used to identify Picea glauca (white
spruce) as the dominant constituent of this forest (Appendices
B). No identifiable macrofossils of P, marianu (black
spruce) were recovered. An open forest is indicated by
macrofossils of Rubus idaeus (wild raspberw), Shepherdia
canadensis (Canada buffaloberry), Fragaria virginiana
(smooth wild
berry), Ro rippa cf. R. rruncora
(blunt-fruited yellow cress), Chenopodium cf. C. berlandieri
(false lamb's-quarters), and Mentha arvensis var. villosa
(field mint), which are all light-demanding shrubs and herbs
(Fig. 4).
Overlying the root and litter layer are 55 cm of dark olive
grey (5Y 312, MunsellTM)silty clay, which constitute the
uppermost part of zone I1 (Fig. 3). These sediments are laminated and contain several large (18-30 cm in diameter,
>1.5 m long) Picea trunks (Fig. 2B, 2C, 3). Other Picea
macrofossils are also found in these fine-grained sediments,
but are less abundant than those from the underlying litter
layer (Fig. 4). One piece of charred wood was recovered
10 cm above the litter layer and is associated with a peak in
charcoal abundance (Fig. 4). The laminated silty clay
(Fig. 5A) and the remains of molluscs and arthropods, including Coleoptera (Appendix A, P1. A-7, fig. 5-9), suggest that a
perennial pond had developed, and that water depth was at
least 1.8 m (e.g. McAndrews et al., 1967) and perhaps more
than 3 m (e.g. Last, 1989).

~acrofossils(# 1 50 ml)

Figure 4. Summary plant macrofossil diagramfor the Andrews site showing the 41 taxa identified. Analysis
by C.H. Yansa. All fossils are seeds or fruits, unless otherwise indicated. These plant macrofossils are illustrated in Appendix A and described in Appendix B.

C.H. Yansa and J.F. Basinger

The macrofossil flora from the pond deposits of zone I1


indicate that the vegetation was species rich (Fig. 4), representing a diversity of microhabitats. Two 3 cm thick organic
layers composed almost entirely of the moss Drepanocladus
polycarpus (Bland. ex Voit) Roth (= Drepanocladus aduncus
(Hedw.) Warnst.) occur near the base of these deposits
(Fig. 4,5A). This sickle-branch moss can either be emergent
or shallowly submerged and is a good indicator of open, wet,
alkaline, calcareous habitats (Crum and Anderson, 1981).
Emergent plants fringing the pond included Epilobium cf.
E. ciliarum (northern fireweed), Typha latifolia (common cattail), Scirpus americanus (three-square bulrush), Carex cf. C.
rostrata (beaked sedge), and Carex cf. C, atherodes (awned
sedge). Aquatic plants included Wolffia arrhiza
(water-meal), Myriophyllum verticillatum var. pectinatum
(whorled water-milfoil), Potamogeton filiformis
(thread-leaved pondweed), P. natans (floating-leaved pondweed), P. obtusifolius (blunt-leaved pondweed), Hippuris
vulgaris (mare's-tail), and Lemna trisulca (ivy-leaved
duckweed). All of these species are indicative of quiet, pH
neutral to alkaline, nutrient-rich waters. The presence of
Rorippa cf. R. truncata, Mentha arvensis var. villosa,
Ranunculus sceleratus (celery-leaved buttercup), and
Chenopodium cf. C. berlandieri indicate that the soils in the
vicinity of the pond were slightly alkaline and nonsaline.
Species such as Geranium cf. G. bicknellii (Bicknell's
geranium), Cirsium cf. C. muticum, and Aster cf.
A, novae-aagliae (New England aster) probably occupied the

Figure 4. (cont.)

pond edges. Juniperus communis (low juniper) was likely to


be found on the more open, well drained knolls or
south-facing slopes. Due to the lack of preserved grasses,
except for one vegetative fragment (Appendix A, P1. A-3,
fig. 8), it is not known if they were a significant component of
this vegetational assemblage.
Near the top of zone 11, the sediments grade from gyttja to
marl. This transition is accompanied by the first appearance
of leaves and buds of Populus balsamifera (balsam poplar)
and P. tremuloides (aspen poplar) and the last appearance of
Picea wood (Fig. 3,5B).

Zone ZZZ Populus, Betula; laminated silty clay


(445-400 cm)
With the exception of the uppermost 5 cm, the marly sediments of zone I11 are laminated. Deposition of this unit
occurred between ca. 10 200 P ed on spruce trunks, which
may be somewhat older than the sediments) and ca. 8800 BP
(Fig. 3; Table 1). The AMS 14Cage of 8800 BP was obtained
from seeds collected 5 cm below the top of this unit; the
uppermost sediments of zone LU are not laminated and contain poorly preserved plant macrofossils. An undulating contact occurs between zone I11 sediments and the overlying
charcoal-rich sandy. clays
- of zone IV (Fig. 2C, 3). These lithological changes suggest a general shallowing of the pond
during deposition of zone 111 sediments. The plant
macrofossils from zone 111 (Fig. 4) also indicate shallowing of

GSC Bulletin 534

Figure 5. A) Cross-section of laminated lacustrine sediments of zone II containing the upper of two
Drepanocladuspolycarpus (sickle-branch)moss layers; samplefrom 480-495 cm; scale bar = 1 cm;
GSC 1999-0440. B) Plan view of leaves of Populus balsamifera (balsam poplar, Pb) and
P. tremuloides (aspenpoplar, Pt) preserved in lacustrine silty clay; sample from 445 cm; scale (on
left end of sample) = 1 cm. GSC 1999-044E. Photographs by C.H.Yansa.

C.H. Yansa and J.F. Basinger

the pond within a parkland setting. Subularia aquatica (water


awl-wort), Eleocharis palustris (creeping spike-rush), and
Typlza latifolia are common constituents of shallow-water
and shoreline environments (Thompson, 1992). Poplar
macrofossils, as well as Betula cf. B. occiderztalis(river birch)
fruits, are common throughout zone 111 sediments (Fig. 4).
The marly sediments indicate a significant increase in dissolved carbonate. Abundant lime-encrusted shoots and
oogonia of Chara sp. (stonewort algae; Appendix A, PI. A-1,
fig. 1,2) indicate that the water was warm, shallow, and carbonate rich. Reduction in the number of Hippuris vulgaris,
Myriophyllurn verticillaturn var. pectinatum, and Lernna
trisulca macrofossils, and increases in Scirpus americanus
seeds and Zannichellia palustris var. palustris (hornedpondweed) fruits indicate a shift in the water chemistry
toward more alkaline and brackish conditions (Fig. 4).
Increased abundance of Chenopodiurn salinurn (saline
goosefoot) throughout zone I11 and the first appearance of
Rurnex maritimus var. fueginus (golden dock) provide evidence of brackish and alkaline soils (Fig. 4). The plant
macrofossils of zone 111, therefore, suggest the development
of alkaline and brackish conditions in a shallowing pond.

Zone ZV - Charcoal; sandy clay (400-390 cm)


ZoneIV consists of a 10 cm thick black (5Y 2.511, MunsellTM)
sandy clay containing abundant charcoal, rootlets, and a few
charred seeds (Fig. 2C, 3,4). The carbonaceous remains were
used to obtain an AMS 14cage of 7670 BP (Table 1). Abundant charcoal indicates that a major fire, or recurring fires,
swept the area during deposition of zone IV sediments. The
occurrence of sand may be attributed to slopewash and/or
deposition of eolian sediment.

drying pond beds, while Rumex imritimus var.fueginus and


Chenopodium salinurn indicate that the soil conditions were
brackish and alkaline.
Increasing fungal and microbial degradation of the fossil
remains near the top of zone V (Appendix A, P1. A-3, fig. 2,
3), and decreasing species diversity and abundance (Fig. 4),
indicate that this wetland was becoming ephemeral to the
point that fossil preservation was deteriorating. The disappearance of fossils correlates with the change from a sandy
clay to a silty clay at a depth of 3.1 m (Fig. 3,4).

INTERPRETATION
The zones of paleoenvironmental change that have just been
described are interpreted below and depicted as a sequence of
phases in Figure 6.

Phase 1 - late glacial


The top 70 cm of zone I sediments were sampled for plant
macrofossils to determine whether a pioneer herbaceous
community could be recognized, for such a vegetation has yet
to be identified from other fossil deposits in the Canadian
Prairie region (Ritchie, 1976,1987). However, the rare fossils
recovered from this deposit are likely reworked. Based on
available evidence, this phase is shown in Figure 6 as a
nonvegetated surface.

Phase 2 - open white spruce forest (ca. 10 200 BP)


Phase 2a - forest development

The dark olive grey (5Y 312, MunsellTM)sandy clay sediments of zone V are nonlaminated, probably due to
bioturbation or wave action (e.g. Last, 1989). Pulmonate gastropods suggest that water depth in zone V was shallower than
in zones I1 (lacustrine phase) and 111(with a branchiate gastropod fauna) (Appendix A, P1. A-7, fig. 5,6; Thompson, 1962;
McAndrews et al., 1967).Macrofossils of the few aquatic and
emergent plant species recovered also indicate that water levels were intermediate between the high-water stand of zones
I1 and I11 and the ephemeral conditions seen today (Fig. 4).
Abundant carbonate nodules and Chara sp, macrofossils
indicate that the water was carbonate rich.

In situ Picea sp, roots (Fig. 2C), which penetrate the 10 cm


thick litter layer (zone 11) into diamicton (Fig. 3), indicate
development of a white spruce forest at the Andrews site by
ca. 10 200 BP. The timing for appearance of this forest at the
Andrews site (Table 1) lags significantly behind the estimated retreat of the Laurentide Ice Sheet from this area at ca.
11 700-1 1 300 BP (Clayton and Moran, 1982). Nevertheless,
based on geomorphological evidence, Clayton (1967) proposed that, on The Missouri Coteau of North Dakota, it took
1000-3000 years after ice sheet retreat for buried stagnant ice
to melt. This interpretation, combined with evidence from the
Andrews site for existence of a well drained spruce forest
habitat at a site that would later become a pond, suggests that
initial development of a forest in this area occurred on a thick
blanket of till overlying stagnant ice, and that the landscape
subsequently experienced inversion (Fig. 6, phase 2a).

Fossils from this zone indicate formation of a


semipermanent pond in a grassland setting. The only emeris known for its capacgent plant of this zone, Typha lat~olia,
ity to produce abundant seeds in disturbed habitats, and to
quickly revegetate when water levels drop (Stewart and
Kantrud, 1972; Thompson, 1992). Fluctuating water levels
may account for the abundance of wet-meadow species well
adapted to disturbance and for an increase in seed production
and species diversity in the surrounding area (Fig. 4).
Ranunculus sceleratus, in particular, is a good indicator of

The Andrews site record is consistent with other studies


from north-central North America that have reported Picea
glauca as a dominant component of early postglacial vegetation (Ritchie and MacDonald, 1986; Ritchie, 1987; Grimm,
1995).Like these other studies, the assemblage reconstructed
for phase 2a at the Andrews site appears to lack a precise
modern analogue. While most taxa of the Andrews site spruce
forest presently are found in the mixed-grass prairie region,
taxa such as Caltha palustris (marsh-marigold) and Cirsium
muticum (swamp thistle) are constituents of the aspen

Zone V - Typha, Ranunculus; nonlaminated sandy


clay (390-310 cm)

GSC Bulletin 534

(< 10 200 BP - 8800 BP)

..............................

C.H. Yansa and J.F. Basinger

parkland and southern boreal forest, while Picea glauca is a


constituent of the Cypress Hills and southern boreal forests.
Climatic conditions supporting P. glauca forests in the southern boreal forest region of Saskatchewan today include lower
mean summer temperature (18OC) and higher mean annual
precipitation (390 mm) than is currently found at the
Andrews site (Ritchie and Harrison, 1993).
Phase 2b - surface collapse
The Picea glauca forest occupied a site that was not initially
water filled, as indicated by the preserved forest soils of lower
zone 11, and the preference of this species for well drained
soils. Eventually, the continued melting of buried stagnant ice
would have resulted in subsidence of the surface until the
water table intersected the base of the newly formed depression, creating a pond at the Andrews site (landscape inversion; Fig. 6, phase 2b). This gradual process of melting of buried ice and collapse of the forest-covered till surface to create
the present knob-and-kettle terrain has been proposed for
similar deposits in North Dakota (Clayton, 1967) and the
Great Lakes region (Warner et al., 1991). A gradual depression and inundation to form a closed basin with ponded water
is supported by the basal, intact soil litters and by the progressive increase in the diversity and abundance of aquatic plant
and animal fossils recovered from the bottom to the top of the
P. glauca-dominated zone I1 (Fig. 4).

Figure 6. Schematic reconstruction of phases 1-6 of


late-glacial to mid-Holocene paleoenvironmental change at
the Andrews site.

LEGEND
White spruce
tree

Mudflat plant

Deciduous tree

Aquatic plant

Shrub

Root

O Cobble

Grasslforb

Emergent plant

c.,=)White spruce

trunk

Phase 2e - pond development


It is unclear from the fossils and lithology whether the spruce
forest at the Andrews site persisted long after pond development. The continued presence of Picea glauca in the lacustrine sediments of zone I1 may have resulted from addition of
long-dead debris and thus predated the pond. Equally plausible is that spruce trunks found buried at various levels within
the laminated lacustrine sedments may be remains of a
drowned forest, the tangle of submerged woody stems resisting decay for centuries before finally being buried. The latter
scenario is more consistent with the uniform 14cages attributed to the woods found below and within the lacustrine sediments of zone 11, as aerial exposure is unlikely to preserve
dead wood for more than decades.

Phase 3 - deciduous parkland (10 200-8800 BP)


From 10 200 to 8800 BP there occurred a change in the flora
at the Andrews site, from open spruce forest (zone 11) to
deciduous parkland (zone 111) dominated by Populus
balsamifera, P. tremuloides, and Betula cf. B. occidentalis
(Fig. 3,4,5B, 6; phase 3). This change probably occurred in
response to regional warming andlor drying, which would
have favoured the growth of broad-leaved angiosperms over
evergreen conifers. Charred wood and abundant charcoal
recovered from the laminated sediments of zone I1 (Fig. 4)
suggest that fire, perhaps associated with increasing aridity,
may have contributed to elimination of spruce. On a regional
scale, Ritchie (1976) attributed the elimination of Picea
glauca from late-glacial environments to increased aridity.
General circulation model (GCM) simulations and pollen
spectra data for north-central North America indicate that
temperature in western Canada was about 2C warmer than
present during the Hypsithermal, a warm and dry interval that
persisted from ca. 10 000-9000 BP until at least 6000 BP and
was associated with summer insolation greater than today
(Ritchie et al., 1983; Kutzbach and Webb, 1991, 1993;
MacDonald, 1993; Vance et al., 1994). Under these conditions, as Szeicz and MacDonald (1994) have shown,
P. glauca is unable to regenerate.
Although evidence indicates that the area became
increasingly arid and water levels declined near the end of
phase 3, deep-water conditions nevertheless persisted
throughout most of this phase, and sufficient moisture was
available to support diverse aquatic, emergent, and
wet-meadow plants within a deciduous broadleaf forest. All
of the taxa recovered from zone 111, except for Aster cf. A.
novae-angliae and Subularia aquatica, occur in the
mixed-grass prairie region today. These two exceptions are
indicative of a cooler and moister climate, although
hummocky morraine has been found by Zoltai (1988) to contribute to microclimate diversity conducive to persistence of
such taxa in an otherwise arid environment.
The transition from laminated to nonlaminated sediments
5 cm below the top of zone I11 indicates shallowing (e.g. Last,
1989). The undulating contact between sediments of zones
I11 and IV (Fig. 2C) is interpreted as a load-contorted contact
between the saturated fine-grained pond deposits of zone I11
and the overlying coarser grained sediments of zone IV, a

GSC Bulletin 534

Table 2. Kettle-fill sites on the n o r t h e r n Great Plains that provide records of early Holocene
revegetation. Only 1 4 c ages obtained from wood and AMS 1 4 C ages obtained from seeds are listed.
Locations of sites that are in the Palliser Triangle are shown in Figure 1.
Site

'C age (a)

Materials dated

Kyle

5053'N, 107"50'W

10 300 + 90 (GSC-5622)'

Beechy

5055'N, 107'40W

10 300 2 90 (GSC-5921)'

Herbert

5022'N, 10715'W

10 200 300 (S-41)


9900 * 350 (S-41)

Crestwynd

4go52'N, 105'39'W

Scrimbit

4g046'N, 105"ll'W

1 Seibold

1 4g055'N, 107"40'W 1

147'03'N, 99"l I'W


Woodworth 47'17'N, 98"53'W
Pond

9750

Tappen

46"46'N, 9g035'W

11 480

Cass

46"46'N, 97"40W

Mclntosh

46"14'N, 9g022W

Val Marie

4g003'N, 107"41'W

* 140 (1-4537)
-

* 300 (W-542)
-

Yansa (1995)

Picea wood

ca. 400

Yansa (1995)

1 wood (unidentified)

4g035'N, 111"49'W

9380 + 110 (TO-1097)'

5 0 ~ 4 ~ N , l 1 0 4 i w10050*llO(AECV-1594C)'
9890

* 120 (AECV-1597C)'

Kupsch (1960)

Ritchie (1966)

Dew (1959), DeVries (1963)

ca. 400

fractionation

feature consistent with shallowing. This is similar to the interpretation presented for the Fletcher site in southeastern
Alberta (Fig. 1, Table 2) by Vickers and Beaudoin (1989). It
is significant that the water table remained sufficiently high
throughout this time and the remainder of the Holocene to
have preserved the underlying fossils and laminated sediments of zones I1 and 111 within a continuously saturated
subsurface environment.

Phase 4 -prairie fires (8800-7700 BP)


It is difficult to ascertain fire frequency during deposition of
the 10 cm thick layer of charcoal-rich sandy clays that make
up zone IV.The rarity of uncharred seeds and plant fragments
may indicate that this deposit represents a single, intense fire.
Accumulation of organic matter during a more humid climatic regime may have made the region vulnerable to conflagration during increasing drought. Alternatively, if a more open
grassland vegetation was present during phase 4, then frequent fires may have consumed most of the litter and left little
uncharred fossil evidence. In either case, phase 4 represents a
significant change in fire regime that is likely attributed to
increasing aridity. Nevertheless, the pond at no time dried out
during this interval.
Removal of surface vegetation as a result of fire apparently resulted in intense slope-wash andlor eolian transport of
sand into the pond. This explanation has been invoked for
occurrence of sandy sediment in kettles elsewhere in
Saskatchewan and in North Dakota (e.g. Moran et al., 1976;
Vreeken, 1994).

Moir (1957)

Picea wood

I1

Malo (1988)

Thompson (1962)
Klassen (1994)
Klassen (1994)

wood (unidentified)
ca. 286
ca. 460
415
same sample
ca. 400

' indicates that I4Caae has been corrected for 8C

DeVries (1963)

Cvancara et al. (1971)

McAndrews et al. (1967)

Cyperaceae seeds
(AMS)
Popu\uswood

4g013'N, 109"lO'W

Fletcher

330

seeds (AMS)
seeds (AMS)

Horseman

Jenner

ca. 400

Salix wood
same sample

Ritchie and DeVries (1964)

Picea wood

9900 + 80 (TO-1711)'
9880 80 (TO-2212)'

*
9500 * 80 (GSC-4098)'

Depth (cm) Reference@)

5020'N, 105"52'W

1 Wrusu

l Location

Hafichuk

Vickers and Beaudoin (1989),


Beaudoin (1992, 1996)
Beaudoin (1992)
Beaudoin (1992)

Beaudoin (1992)

Phase 5 - semipermanent prairie pond


(ca. 7700-5800 BP)

The plant macrofossils, gastropods, and sedimentology indicate that water levels during phase 5 were intermediate
between the highstand that occurred during phases 2c and 3
and the ephemeral conditions seen today. The depth of the
Andrews pond is interpreted to decrease continually through
phases 3, 4, and 5 . Water levels were sufficiently high
throughout phase 5 to sustain a semipermanent prairie pond
vegetation. At no time during deposition of zone V sediments
did the basin dry long enough for plant remains in the sediments to decay or oxidize, and there is no evidence for the formation of a soil.
Precipitation during phase 5 may have been variable,
because the Andrews site plant macrofossil record suggests
that water levels may have fluctuated (Fig. 4). On the basis of
comparison with modern semipermanent prairie ponds
(Stewart and Kantrud, 1972), it is proposed that a mudflat
environment expanded from the shoreline toward the centre
of the basin whenever the water table lowered, whereas the
fringe of cattails would expand at the expense of mud flats
when the water level rose. Eventually, the water level fell low
enough, and the sediment surface was exposed long enough,
to prevent plant fossil preservation.

C.H. Yansa and J.F. Basinger

Phase 6 - ephemeral pond


About 5800 BP, the Andrews site pond became ephemeral
and preservation of macrofossils was precluded (Fig. 6,
phase 6). Paleoclimate reconstructions indicate that a climate
more arid than at present prevailed in western Canada at that
time (Vance et al., 1994). While aridity likely contributed to
the cessation of fossil preservation at the Andrews site, additional factors may have been involved. It may be that, by the
mid-Holocene, the Andrews site and perhaps most other kettles within the region were filled with enough sediment to
reduce the water-holding capacity of these basins, and thus
restrict fossil preservation to below the water table. As well,
fluctuations in local groundwater supply may have altered the
water table at this and other sites in this region (e.g. Last and
Slezak, 1986), thus limiting the extent to which fossils could
be preserved. Therefore, climate, sedimentological processes, and local hydrology may all have contributed to the
discontinuation of the Andrews site fossil record at ca.
5800 BP.

CONCLUSIONS
The Andrews site provides one of the most complete records
of earliest postglacial vegetation currently available in the
northern Great Plains region, including remains of an in situ
Picea glauca forest litter and underlying diamicton. The
record of the Andrews site, in the context of other sites known
from the northern Great Plains, contributes to local and
regional paleoenvironmental interpretation.
Picea glauca forests occupied North Dakota at
ca. 12 000 BP, and extended into southern Saskatchewan by
ca. 10 200 BP at the Andrews site and about the same time
(ca. 10 300 BP) at the Kyle and Beechy sites (Fig. 1; Table 2;
Moir, 1957; Klassen, 1994; Grimm, 1995). To the southwest,
open grassland has apparently existed in Montana over the
past 12 000 years, with spruce being conspicuously absent
(Barnosky, 1989).In southeastern Alberta, Populus was present, but Picea apparently absent (Table 2; Beaudoin, 1992,
1996).
On The Missouri Coteau throughout North Dakota and
southern Saskatchewan, spruce forests are interpreted to have
been growing on a blanket of till overlying stagnant ice (Clayton,
1967; Yansa, 1998). Melting of buried ice caused collapse of
the land surface to form the knob-and-kettle terrain that characterizes this upland (Clayton, 1967; Yansa, 1998). Spruce
forests likely persisted on the slopes and knolls surrounding
the ponds.
The spruce forests in northern North Dakota were apparently replaced by grassland vegetation at ca. 10 000 BP
(Grirnm, 1995). Data from the Andrews site indicates that
spruce forests in southern Saskatchewan were succeeded by
deciduous parklands prior to establishment of open prairie.
Balsam poplar, aspen poplar, and river birch became established at some time after 10 200 BP, and persisted until at least
8800 BP.

The moist climate and deep ponds of the Andrews area are
consistent with early Holocene records reported for a number
of kettles and lakes in southern Saskatchewan and southern
Alberta (Last, 1990; Klassen, 1994; Beaudoin, 1996).
Beaudoin et al. (1996) suggested that high water levels
throughout the region at that time resulted from the influx of
meltwater into aquifers from the melting of buried stagnant
ice. Such a source of water may have buffered the Iocal
environments against increasing warmth and dryness of the
early Holocene Hypsithermal, and contributed to the apparent delay in the effects of early Holocene aridity in this
region.
A floral shift toward plants tolerant of shallow, alkaline,
and brackish water occurred in the upper part of zone 111at the
Andrews site (Fig. 6, phase 3), and has also been noted in
other kettle-fill and lake studies from the region (Sauchyn and
Sauchyn, 1991; Beaudoin, 1992; Klassen, 1994; Wilson
et al., 1997). The increase in aridity, identified in phase 3,
intensified during phase 4 at the Andrews site, indicating the
onset of Hypsithermal effects at the Andrews site sometime
between ca. 8800 and 7700 BP. Delay in the appearance of
peak Holocene aridity until the mid-Holocene has also been
reported from the American Midwest, where it has been identified as time transgressive and corresponding to the retreat of
the Laurentide Ice Sheet across north-central North America
(Wright, 1976; Winkler et al., 1986; Baker et al., 1992; Dean
et al., 1996). Evidence from the Mackenzie Delta, central
Alberta, and Montana, however, reveals that dry lake basins
and drought-tolerant taxa were common from 10 000 to
8000BP (Ritchie et al., 1983; Barnosky, 1989; Schweger and
Hickman, 1989). Thus, in addition to protracted release of
water from subsurface ice, the delay in appearance of
Hypsithermal drought at the Andrews site could be attributed
to a lag in deglaciation in this region.
Water levels in the Andrews site pond appear to have continually decreased, with minor fluctuations, through to about
5800 BP, when the pond become ephemeral and preservation
of macrofossils ceased. Reconstruction of an interval of fluctuating water levels from ca. 7700 to 5800 BP, within a
long-term climatic drying trend, agrees with the interpretation by Vance et al. (1992), based on the Chappice Lake
record of southeastern Alberta, of brief moist intervals occurring during a largely arid period between 7300 and 6000 BP.
Vance et al. (1992) suggested that the regional climate
between 6000 and 4000 BP was more arid than at present,
which is consistent with the termination of the Andrews site
fossil record at about 5800 BP. Although alacustrine environment persisted throughout the Holocene in some basins, such
as ChappiceLake (Vance et al., 1992,1993), the vast majority
of ponds in the region dried and vanished, as was the case at
the Andrews site.

ACKNOWLEDGMENTS
We thank A. Andrews for permission to collect samples on
her property; L. Fritz and S.J. Yansa for field assistance;
V.H. Harms and P.A. Ryan for permitting preparation of a
comparative collection from specimens housed in the Fraser

GSC Bulletin 534


Herbarium; the Geological Survey of Canada and the Alberta
Environmental Centre-Vegreville for radiocarbon ages; and
A.E. Aitken, B.A. LePage, A.K. Burt, C. Greenwood, and
M.J. Reeves for technical advice and assistance. This manuscript has benefited considerably from critical comments
made by B.A. LePage, D.S. Lemmen, R.W. Mathewes,
R.J. Mott, and R.E. Vance. Funding was provided by the
Palliser Triangle Global Change Project (to J.F. Basinger,
A.E. Aitken, and C.H. Yansa), Geological Survey of Canada;
and by the Natural Sciences and Engineering Research Council of Canada (NSERC operating grant #OGP0001334 to
J.F. Basinger).

REFERENCES
Acton, D.F., Clayton, J.S., Ellis, J.G., Christiansen, E.A.,
and Kupsch, W.O.
1960: Physiographic divisions of Saskatchewan;Saskatchewan Research
Council, Geology Division, no. 1, scale 1:l 520 640.
Baker, R.G., Maher, L.J., Chumbley, C.A., and Van Zant, K.L.
1992: Patterns of Holocene environmental change in the Midwestern
United States; Quaternary Research, v. 37, p. 379-389.
Barker, W.T. and Whitman, W.C.
1988: Vegetation of the northern Great Plains; Rangelands, v. 10,
p. 266-272.
Barnosky, C.W.
1989: Postglacial vegetation and climate in the northwestern Great Plains
of Montana; Quaternary Research, v. 31, p. 57-73.
Barnosky, C.W., Grimm, E.C., and Wright, H.E., Jr.
1987: Towards a postglacial history of the northern Great Plains: a review
of the paleoecologic problems; Annals of Carnegie Museum, v. 56,
p. 259-273.
Bateman, R.M.
1991: Palaeoecology; in Plant Fossils in Geological Investigation: the
Palaeozoic,(ed.) C.J. Cleal; Ellis Horwood, New York, p. 34-1 16.
Beaudoin, A.B.
1992: Early Holocene paleoenvironmental data preserved in
'non-traditional' sites; Palliser Triangle Global Change Meeting,
Regina, Saskatchewan, Program and Abstracts, p. 1-2.
1996: An early Holocene macrofossil record from the Fletcher site
(DjOw-1) in southern Alberta; Canadian Association of Geographers Meeting, Saskatoon, Saskatchewan, Program and Abstracts,
p. 61-62.
Beaudoin, A.B., Yansa, C.H., and Vance, R.E.
1996: A model for late glacial-early Holocene landscape development on
the Northern Plains; Canadian Association of Geographers Meeting, Saskatoon, Saskatchewan, Program and Abstracts, p. 62-63.
Birks, H.J.B. and Birks, H.H.
1980: Quaternary Palaeoecology; University Park Press, Baltimore,
Maryland, 289 p.
Christiansen, E.A.
1961: Geology and ground-water resources of the Regina area,
Saskatchewan; Saskatchewan Department of Mineral Resources,
Report 2.72 p.
Clayton, L.
1967: Stagnant-glacierfeatures of The Missouri Coteau in North Dakota;
in Glacial Geology of The Missouri Coteau, (ed.) L. Clayton and
T.F. Freers; North Dakota Geological Survey, Miscellaneous
Series, v. 30, p. 25-46.
Clayton, L. and Moran, S.R.
1982: Chronology of Late Wisconsinan glaciation in middle North
America; Quaternary Science Reviews, v. 1, p. 55-82.
Crum, H.A. and Anderson, L.E.
1981: Mosses of Eastern North America, Volumes 1 and 2; Columbia University Press, New York, 1328 p.
Cvancara, A.M., Clayton, L., Bickley, W.B., Jr., Jacobs, A.F.,
Ashworth, A.C., Brophy, J.A., Shay, C.T., Delorme, L.D.,
and Lammers, G.E.
1971: Paleolimnology of late Quaternary deposits: Seibold site, North
Dakota; Science, v. 171, p. 172-174.

Dean, W.E., Ahlbrandt, T.S., Anderson, R.Y., and Bradbury, J.P.


1996: Regional aridity in North America during the middle Holocene;The
Holocene, v. 6, p. 145-155.
DeVries, B.
1963: An investigation of a late-glacial deposit frornThe Missouri Coteau
in Saskatchewan;M.Sc, thesis, University of Manitoba, Winnipeg,
Manitoba, 88 p.
Dew, J.
1959: Post-glacial forest at Kayville uncovered; The Blue Jay
(Saskatchewan Natural History Newsletter), v. 17, p. 20-21.
Environment Canada
1993: Canadian Climate Normals, 1961-1990; Environment Canada,
Atmospheric Environment Service, Ottawa (CD-ROM).
Faegri, K., Kaland, P., and Krzywinski, K.
1989: Textbook of Pollen Analysis; John Wiley & Sons, Chichester,
United Kingdom, 328 p. (fourth edition).
Gale, S.J. and Hoare, P.G.
1991: Quaternary Sediments: Petrographic Methods for the Study of
Unlithified Rocks; Halstead Press, Toronto, Ontario, 323 p.
Grimm. E.C.
1992: ilia^^ 1.12 and Tilia.graphTM1.19; lllinois State Museum,
Research and Collection Center, Springfield, Illinois.
1995: Recent palynological studies from lakes in the Dakotas; Geological
Society of America, North-Central Section-South-Central Section
Meeting, Lincoln, Nebraska, Program and Abstracts, p. 54.
Gullett, D.W. and Skinner, W.R.
1992: The state of Canada's climate: temperature change in Canada
1895-1991; Environment Canada, Atmospheric Environment
Service, State of the Environment Report No. 99-2.36 p.
Harms, V.L., Hudson, J.H., and Ledingham, G.F.
1986: Rorippa truncata, the blunt-fruited yellow cress, new for Canada,
and R. tenerrima, the slender yellow cress, in southern
Saskatchewan and Alberta; Canadian Field-Naturalist,. v. 100,
p. 45-51.
Hubbard, D.E. and Linder, R.L.
1986: Spring runoff retention in prairie pothole wetlands; Journal of Soil
and Water Conservation, v. 41, p. 122-125.
Johnson, J.H.
1961: Limestone-buildingAlgae and Algal Limestones; Colorado School
of Mines, Department of Publications, Golden, Colorado, 297 p.
Klassen, R.W.
1992: Nature, origin, and age relationships of landscape complexes in
southwestern Saskatchewan; GBographie physique et Quaternaire,
v. 46, p. 361-388.
1994: Late Wisconsinan and Holocene history of southwestern
Saskatchewan; Canadian Journal of Earth Sciences, v. 31,
p. 1822-1837.
Kupsch, W.O.
1960: Radiocarbon-dated organic sediment near Herbert, Saskatchewan;
American Journal of Science, v. 258, p. 282-292.
Kutzbach, J.E. and Webb, T., In.
1991: Late Quaternary climatic and vegetational change in eastern North
America: concepts, models, and data; in Quaternary Landscapes,
(ed.) L.C.K. Shane and E.J. Cushing; University of Minnesota
Press, Minneapolis, Minnesota, p. 175-217.
1993: Concephlal basis for understanding late Quaternary climates; in
Global Climates Since the Last Glacial Maximum, (ed.)
H.E. Wright, Jr., J.E. Kutzbach, T. Webb, HI, W.F. Ruddiman,
F.A. Street-Perrott, and P.J. Bartlein; University of Minnesota
Press, Minneapolis, Minnesota, p. 5-1 1.
Last, W.M.
1989: Continental brines and evaporites of the northern Great Plains of
Canada; Sedimentary Geology, v. 64, p. 207-221.
1990: Paleochemistry and paleohydrology of Ceylon Lake, a
salt-dominated playa basin in the northern Great Plains, Canada;
Journal of Paleolimnology, v. 4, p. 219-238.
1994: Paleohydrology of playas in the northern Great Plains: perspectives
from Palliser's Triangle; in Paleoclimate and Basin Evolution of
Playa Systems, (ed.) M.R. Rosen; Geological Society of America,
Special Paper 289, p. 69-80.
Last, W.M. and Slezak, L.A.
1986: Paleohydrology, sedimentology, and geochemistry of two
meromictic saline lakes in southern Saskatchewan; Gkographie
physique et Quaternaire, v. 40, p. 5-15.
1988: The salt lakes of western Canada: a paleolimnological overview;
Hydrobiologia, v. 158, p. 301-316.

C.H. Yansa and J.F. Basinger


Lemmen, D.S., Dyke, L.D., and Edlund, S.A.
1993: 'The Geological Survey of Canada's Integrated Research and Monitoring Area (1RMA) projects: a contribution to Canadian global
change research; Journal of Paleolimnology, v. 9, p. 77-83.
LePage, B.A. and Basinger, J.F.
1995: Evolutionary history of the genus Pseudolarix Gordan (Pinaceae);
International Journal of Plant Science, v. 156, p. 910-950.
Levesque, P.E.M., Dinel, H., and Larouche, A.
1988: Guide to the Identification of plant macrofossils in Canadian
peatlands; Agriculturc Canada, Research Branch, Publication
No. 1817,65 p.
Lissey, A.
1971: Depression-focused transient groundwater flow patterns in
~ a n i t o b ain
; Geoscience Studies in Manitoba, (ed.) A.C. 'l'urnock;
Geoloaical Association of Canada. Soecial Paper No. 9. o. 333-34 1.
Looman, J. an'h Best, K.F.
1987: Budd's flora of the Canadian Prairie Provinces; Agriculture
Canada, Research Branch, Publication No. 1662, 863 p. (revised
edition).
MacDonald, G.M.
1993: Methodological falsification and the interpretation of palaeoecological records: the cause of the early Holocene birch decline in
western Canada; Review of Palaeobotany and Palynology, v. 79,
p. 83-97.
Malo, D.D.
1988: The Holocene sedimentation environment in a seasonal wetland in
eastern North Dakota; Soil Survey Horizons, v. 10, p. 141-152.
Martin, A.C. and Barkley, W.D.
1961: Seed Identification Manual; University of California Press,
Berkeley, California, 221 p.
McAndrews, J.H., Stewart, R.E., Jr., and Bright, R.C.
1967: Paleoecology of a prairie pothole: a preliminary report; in Glacial
Geology of The Missouri Coteau, (ed.) L. Clayton and T.F. Freers;
North Dakota Geological Survey, Miscellaneous Series, v. 30,
p. 101-1 13.
Moir, D.R.
1957: An occurrence of buried coniferous wood in the Altamont Moraine
in North Dakota; Proceedings of North Dakota Academy of Science, v. 11, p. 31-36.
Montgomery, F.H.
1977: Seeds and Fruits of Plants of Eastern Canada and Northeastern
United States; Universityof TorontoPress, Toronto, Ontario, 232 p.
Moran, S.R., Arndt, M., Bluemle, J.P., Camara, M., Clayton, L.,
Fenton, M.M., Harris, K.L., Hobbs, H.C., Keatinge, R.,
Sackreiter, D.K., Salomon, N.L., and Teller, J.
1976: Quaternary stratigraphy and history of North Dakota, southern
Manitoba, and northwestern Minnesota; in Quaternary stratigraphy
of North America, (ed.) W.C. Mahaney; Dowden, Hutchinson and
Ross, Inc., Stroudsburg, Pennsylvania, p. 133-158.
Mott, R.J.
1973: Palynological studies in central Saskatchewan: pollen stratigraphy
from lake sediment sequences; Geological Survey of Canada, Paper
72-49, 18 p.
Mott, R.J. and Christiansen, E.A.
1981: Palynological study of slough sediments from central
Saskatchewan; in Current Research, Part B; Geological Survey of
Canada, Paper 81-lB, p. 133-136.
Risser, P.G., Birney, E.C., Blocker, H.D., May, S.W., Parton, W.J.,
and Weins, J.A.
1981: The True Prairie Ecosystem; Hutchinson Ross Publishing Co.,
Stroudsburg, Pennsylvania, 557 p.
Ritchie, J.C.
1966: Aspects of the late-Pleistocenehistory of the Canadian flora; in The
Evolution of Canada's Flora, (ed.) R.L. Taylor and R.A. Ludwig;
University of Toronto Press, Toronto, Ontario, p. 66-80.
1976: The late-Quaternary vegetational history of the western interior of
Canada; Canadian Journal of Botany, v. 54, p. 1793-1818.
1987: Postglacial Vegetation of Canada; Cambridge University Press,
Cambridge, United Kingdom, 178 p.
Ritchie, J.C. and DeVries, B.
1964: Contributions to the Holocene paleoecology of west-central
Canada; Canadian Journal of Botany, v. 42, p. 677-692.

Ritchie, J.C. and Harrison, S.P.


1993: Vegetation, lake levels, and climate in western Canada during the
Holocene; in Global Climates Since the Last Glacial Maximum,
(ed.) H.E. Wright, Jr., J.E. Kutzbach,T. Webb, III,W.F. Ruddiman,
F.A. Street-Perrott, and P.J. Bartlein; University of Minnesota
Press, Minneapolis, Minnesota, p. 401414.
Ritchie, J.C. and MacDonald, G.M.
1986: The patterns of post-glacial spread of white spruce; Journal of
Biogeography, v. 13, p. 527-540.
Ritchie, J.C., Cwynar, L.C., and Spear, R.W.
1983: Evidence from north-west Canada for an early Holocene
Milankovitch thermal maximum; Nature, v. 305, p. 126-128.
Sauchyn, M.A. and Sauchyn, D.J.
1991: A continuous record of Holocene pollen from Harris Lake, southwestern Saskatchewan, Canada; Palaeogeography,
Palaeoclimatology,Palaeoecology, v. 88, p. 13-23.
Schweger, C.E. and Hickman, M.
1989: Holocene paleohydrology of central Alberta: testing the general-circulation-model climate simulations; Canadian Journal of
Earth Sciences, v. 26, p. 1826-1833.
Stewart, R.E. and Kantrud, H.A.
1972: Vegetation of prairie potholes, North Dakota, in relation to quality
of water and other environmental factors; United States Geological
Survey, Professional Paper 585-D, 36 p.
Szeicz, J.M. and MacDonald, G.M.
1994: Age-dependent tree-ring growth responses of subarctic white
spruce to climate; Canadian Journal of Forest Research, v. 24,
p. 120-132.
Thompson, G.G.
1962: Postglacial fresh-water limestone, marl and peat from south-central
North Dakota; Proceedings of the North Dakota Academy of Science, v. 16, p. 16-22.
Thompson, K.
1992: The functional ecology of seed banks; in Seeds: The Ecology of
Regeneration in Plant Communities. led.) M. Fenner: C.A.B. International, Wallingford, United Kingdom, p. 231-258.
Vance, R.E.
1997: The Geological Survey of Canada's Palliser Triangle Global
Change Project: a multidisciplinary geolimnological approach to
predicting potential global change impacts on the northern Great
Plains; Journal of Paleolimnology, v. 17, p. 3-8.
Vance, R.E. and Last, W.M.
1994: Paleolimnology and global change on the southern Canadian Prairies; in Current Research 1994-B; Geological Survey of Canada,
p. 49-58.
Vance, R.E. and Mathewes, R.W.
1994: Deposition of modern pollen and plant macroremains in a
hypersaline prairie lake basin; Canadian Journal of Botany, v. 72,
p. 539-548.
Vance, R.E., Beaudoin, A.B., and Luckman, B.H.
1994: The paleoecological record of 6 ka BP climate in the Canadian Prairieprovinces; GBographiephysiqueetQuaternaire,v. 49, p. 8 1-98.
Vance, R.E., Clague, J.J., and Mathewes, R.W.
1993: Holocene paleohydrology of a hypersaline lake in southeastern
Alberta; Journal of Paleolimnology, v. 8, p. 103-120.
Vance, R.E., Mathewes, R.W., and Clague, J.J.
1992: 7000 year record of lake-level change on the northern Great Plains:
a high-resolution proxy ofpast climate; Geology, v. 20, p. 879-882.
Vickers, J.R. and Beaudoin, A.B.
1989: A limiting AMS date for the Cody Complex occupation at the
Fletcher site, Alberta, Canada; Plains Anthropologist, v. 34,
p. 261-264.
Vreeken, W.J.
1994: A Holocene soil-geomorphic record from the Ham site near Frontier, southwestern Saskatchewan; Canadian Journal of Earth Sciences, v. 31, p. 532-543.
Warner, B.G.
1990: Plant macrofossils; in Methods in Quaternary Ecology, (ed.)
B.G. Warner; Geological Association of Canada, Geoscience
Canada Reprint Series 5, p. 53-64.
Warner, B.G., Kubiw, H.J., and Karrow, P.F.
1991: Origin of a postglacial kettle-fill sequence near Georgetown,
Ontario; Canadian Journal of Earth Sciences, v. 28, p. 1965-1974.

GSC Bulletin 534


Wilson, S.E., Smol, J.P., and Sauchyn, D.J.
1997: A Holocene paleosalinity diatom record from southwestern
Saskatchewan, Canada: Harris Lake revisited; Journal of
Paleolimnology, v. 17, p. 23-31.
Winkler, M., Swain, A.M., and Kutzbach, J.E.
1986: Middle Holocene dry period in the northern midwestern United
States: lake levels and pollen stratigraphy; Quaternary Research,
v. 25, p. 235-250.
Winter, T.C.
1989: Hydrologic studies of wetlands in the northern prairie; in Northern
Prairie Wetlands, (ed.) A.G. van der Valk; Iowa State Press, Ames,
Iowa, p. 16-54.

Wright, H.E.,Jr.
1976: The dynamic nature of Holocene vegetation: a problem in
paleoclimatology, biogeography, and stratigraphic nomenclature;
Quaternary Research, v. 6, p. 581-596.
Yansa, C.H.
1995: An early postglacial record of vegetation change in southern
Saskatchewan, Canada; M.Sc. thesis, University of Saskatchewan,
Saskatoon, Saskatchewan, 273 p.
1998: Holocene paleovegetation and paleohydrology of a prairie pothole
in southern Saskatchewan, Canada; Journal of Paleolimnology,
v. 19, p. 429441.
Zoltai, S.C.
1988: Wetland environment and classification; in Wetlands of Canada;
National Wetlands Working Group. Ecological Land Classification
Series No. 24, Supply and Services Canada, Ottawa, Ontario,
p. 1-26.

C.H. Yansa and J.F. Basinger

APPENDIX A

This appendix contains SEM micrographs of the 41 plant macrofossil taxa, plus associated animal fossils,
identified at the Andrews site. See Appendix B for descriptions and remarks.

GSC Bulletin 534

PLATE A-1
Figure 1. Chara sp. (stonewort algae) oogonium. Scale bar = 0.5 mm.
Figure 2. Chara sp. shoot fragments. Scale bar = 1.0 mm.
Figure 3. Picea glauca (white spruce) cone. Abundant seeds (as in fig. 4) were contained in cone.
Scale bar = 5.0 mm.
Figure 4. Picea glauca winged seed. Scale bar = 1.0 mm.
Figure 5. Picea glauca needle. Scale bar = 1.0 mm.
Figure 6. Juniperus communis (low juniper) leaf. Scale bar = 1.0 mm.
Figure 7. Typha lafifolia (common cattail) achene. Scale bar = 0.5 mm.
Figure 8. Potamogeton filiformis (thread-leaved pondweed) drupe. Scale bar = 0.5 mm.
Figure 9. Potamogeton natans (common floating-leaved pondweed) drupe. Scale bar = 0.5 mm.

C.H. Yansa and J.F. Basinger

Plate A-1

GSC Bulletin 534

PLATE A-2
Figure 1. Modern Picea glauca (white spruce) inner cuticle showing anticlinal walls. Scale bar = 50 pm.
Figure 2. Modern Picea glauca outer cuticle showing stomatal complexes. Scale bar = 0.1 mm.
Figure3. Modern Picea mariana (black spruce) inner cuticle showing anticlinal walls. Scale bar=50 p'~.
Figure 4. Modern Picea mariana outer cuticle showing stomatal complexes. Scale bar = 0.1 mm.
Figure 5. Fossil Picea glauca inner cuticle showing anticlinal ridges. Scale bar = 50 pn.
Figure 6. Fossil Picea glauca outer cuticle showing stomatal complexes. Scale bar = 0.1 mm.

C.H. Yansa and J.F. Basinger

Plate A-2

GSC Bulletin 534

PLATE A-3
Figure 1. Cross-sections of Potamogeton filiformis (thread-leaved pondweed) drupes: fossil (left); and
modern (right). Scale bar = 1.0 mm.
Figure 2. Fungal and microbial destruction of Potamogeton sp. drupes in zone V, where more testas are
also found. Scale bar = 1.0 mm.
Figure3. Close-up (at arrow in fig. 2) of Potamogetonsp.drupe, showing in situ fungi. Scale bar = 0.5 pm.
Figure 4. Potamogeton obtusifolius (blunt-leaved pondweed) drupe. Scale bar 0.5 mm.
Figure 5. Potamogeton pectinatus (sago or fennel-leaved pondweed) drupe. Arrow at persistent
attachment of drupe, which is a feature characteristic of this species. Scale bar = 0.5 mm.
Figure 6. Potamogeton vaginatus (sheathed pondweed) drupe. Scale bar = 0.5 mm.
Figure 7. Zannichellia palustris var, palustris (horned pondweed) drupe. Scale bar = 0.5 mm.
Figure 8. Gramineae (grass) glumes and awn. Scale bar = 0.5 mm.

C.H. Yansa and J.F. Basinger

Plate A-3

GSC Bulletin 534

PLATE A-4
Figure 1. Carexcf. C. atherodes (awned sedge) achene. Scale bar = 0.5 mm.
Figure 2. Carexcf. C. rostrata (beaked sedge) achene. Scale bar = 0.5 mm.
Figure 3. Eleocharis palustris (creeping spike-rush) achene. Scale bar = 0.5 mm.
Figure 4. Scirpus americanus (three-square bulrush) achene. Scale bar = 0.5 mm.
Figure 5. Scirpus cf. S. validus (great bulrush) achene with basal bristles. Scale bar = 0.5 mm.
Figure 6. Lemna trisulca (ivy-leaved duckweed) thallus. Scale bar = 1.0 mm.
Figure 7. Wolffia arrhiza (water-meal) thallus. Scale bar = 0.5 mm.
Figure 8. Populus balsamifera (balsam poplar) bud scale. Scale bar = 1.0 mm.
Figure 9. Populus tremuloides (aspen poplar) bud scale. Scale bar = 0.5 mm.

C.H. Yansa and J.F. Basinger

Plate A-4

GSC Bulletin 534

Figure 1. Betula cf. B. occidentalis (river birch) samara (wings not preserved). Scale bar = 0.5 mm.
Figure 2. Rumex maritimus var. fueginus (golden dock) achene. Scale bar = 0.5 mm.
Figure 3. Rumex maritimus var. fueginus perianth segments. Scale bar = 0.5 mm.
Figure 4. Chenopodium cf. C. berlandieri (false lamb's-quarters) seed. Scale bar = 0.5 mm.
Figure 5. Chenopodium salinum (saline goosefoot) seed. Scale bar = 0.5 mm.
Figure 6. Caltha palustris (marsh-marigold)seed. Scale bar = 0.5 mm.
Figure 7. Ranunculus sceleratus (celery-leaved buttercup) achene. Scale bar = 0.5 mm.
Figure 8. Rorippa cf. R. truncata (blunt-fruited yellow cress) seed. Scale bar = 0.5 mm.
Figure 9. Subularia aquatica (water awl-wort) seed. Scale bar = 0.5 mm.

C.H. Yansa and J.F. Basinger

Plate A-5

GSC Bulletin 534

PLATE A-6
Figure 1. Fragaria virginiana (smooth wild strawberry) achene. Scale bar = 0.1 mm.
Figure 2. Lemna sp. (Duckweed) achene. Scale bar = 0.1 mm.
Figure 3. Rubus idaeus (wild red raspberry) endocarp. Scale bar = 0.5 mm.
Figure 4. Geranium cf. G. bicknellii (Bicknell's geranium) seed. Scale bar = 0.5 mm.
Figure 5. Shepherdia canadensis (Canada buffaloberry; soapberry) achene. Scale bar = 0.5 mm.
Figure 6. Epilobium cf. E. ciliatum (northern fireweed) seed. Scale bar = 0.5 mm.
Figure 7. Hippuris vulgaris (mare's tail) seed. Scale bar = 0.5 mm.
Figure 8. Myriophyllum sp. (water-milfoil) mericarp. Scale bar = 0.5 mm.
Figure 9. Myriophyllum verticillatum var. pectinaturn (whorled water-milfoil) bract. Scale bar = 0.5 mm.

C.H. Yansa and J.F. Basinger

Plate A-6

GSC Bulletin 534

PLATE A-7
Figure 1. Lycopus americanus (water-horehound)nutlet. Scale bar = 0.5 mm.
Figure 2. Mentha anlensis var. villosa (field mint) nutlet. Scale bar = 0.5 mm.
Figure 3. Astercf. A. novae-angliae (New England aster) achene. Scale bar = 0.5 mm.
Figure 4. Cirsium cf. C. muticum (swamp thistle) achene. Scale bar = 0.5 mm.
Figure 5. Gastropoda (branchiate gastropod) shell. Scale bar = 0.5 mm.
Figure 6.Gastropoda (pulmonate gastropod) shell. Scale bar = 0.5 mm.
Figure 7. Pisidium sp. (clam) shell. Scale bar = 0.5 mm.
Figure 8. Coleoptera (beetle) elytron. Scale bar = 0.5 mm.
Figure 9. Arthropoda (insect), unidentified. Scale bar = 0.5 mm.

C.H. Yansa and J.F. Basinger

Plate A-7

GSC Bulletin 534

APPENDIX B
This appendix provides descriptions and remarks on the 41 plant macrofossil taxa, plus associated animal
fossils, identified at the Andrews site. See Appendix A for illustrations.
Species

Figure(s), Appendix A

Description and remarks

Chara sp. L. (stonewort


algae)

PI. A-1, fig. 1, 2

.3 mm in diameter, with helicoid ridges. Shoots up to 4.4 mm


Elliptical oogonia, 0.9-1.1 mm long, 1.&I
long and 0.8-1.0 mm in diameter. Both oogonia and shoots may be encrusted with CaC03.

Picea glauca (Moench)


Voss (white spruce)

PI. A-1, fig. 3-5; PI. A-2,


fig. 1,2, 5, 6

Seed cones, 27.0-41.0 mm long, 11.&15.0 mm wide. Seeds, 3.2-3.4 mm long, 1.7-1.8 mm wide.
Features of cones, scales, and seeds fall within range of living Picea glauca, and differ from those of F!
mariana (Miller) B.S. & P (black spruce). External morphological features of needles cannot be used to
distinguish between these two species; however, cuticle micromorphologyIs diagnostic. Inner cuticle of
fossil needles (PI. A-3, fig. 1) compares favourably to that of modern F! glauca (PI. A-2, fig. 3) and differs
from that of F! mariana (PI. A-2, fig. 5). Wood of Picea is not identifiable to species, but presumed to be
that of P glauca based on the association.

Juniperus communis L.
(low juniper)

PI. A-1, fig. 6

Leaves narrowly awl shaped, subulate, apex acute, base truncate and jointed, 5.0-5.6 mm long,
0.8-0.9 mm wide, 0.7-0.9 mm thick.

Typha latifolia L.
(common cattail)

PI. A-I, fig. 7

Achenes minute, 1.O-1 .Imm long, 0.30-0.33 mm wide, 0.27-0.30 mm thick; ellipsoidal, apex narrow,
truncate with the micropylar operculum appearing mucronate. Seed coat has a finely areolate texture.

Potamogeton filiformis
Pers. (thread-leaved
pondweed)

PI. A-1, fig. 8; PI. A-3,


fig. 1

Drupe obovoid, asymmetrical, 2.4-2.7 mm long, 1.5-1.6 mm wide, 1.0-1.1 mm thick. Style short, up to
0.4 mm long. Dorsal germination valve extending from base and nearly reaching the style, lateral faces of
drupe flattened with central depression. Drupe single-seeded, seed curved around prominent condyle,
embryo campylotropous (PI. A-3, fig. 3). Seed coat is smooth. Drupes of this species are significantly
smaller than those of other species of this genus.

Potamogeton natans L.
(floating-leaved
pondweed)

PI. A-1, fig. 9

.2 mm thick.
Drupes obovoid to obliquely elliptic, asymmetrical, 3.3-3.6 mm long, 2.0-2.4 mm wide, 1.&I
Dorsal margin rounded with low spines, lateral faces flattened with central depression. Dorsal germination
valve extending from base and nearly reaching the style. Drupe single seeded, seed strongly hook shaped
and curved around prominent condyle, embryo campylotropous. Seed coat areolate.

Potamogeton
obtusifolius Mert. &
Koch. (blunt-leaved
pondweed)

PI. A-3, fig. 4

Drupes obovate, asymmetrical, 3.4-3.6 mm long, 2.1-2.4 mm wide, 0.9-1.2 mm thick. Dorsal margin
prominent with sharply corrugated edge, lateral faces flattened with pronounced external and central
depression. Germination valve extending entire length of fruit. Drupe single seeded. Seed coat areolate.

Potamogeton
pectinatus L. (sago
pondweed or fennelleaved)

PI. A-3, fig. 5

Drupes obliquely obovate to nearly circular, asymmetrical, 3.5-3.8 mm long, 2.5-3.1 mm wide,
1.6-1.9 mm thick. Germination valve extends only one-half to two-thirds the length of the rounded dorsal
margin and is not in contact with the slender style. Drupe single seeded, seed U-shaped and curved
around prominent condyle, embryo campylotropous.Fossil drupes are superficially comparable to those of
Potamogeton vaginatusTurcz. achenes (PI. A-3, fig. 8), but differ in that they are more circular and bear a
persistent attachment (PI. A-3, fig. 7).

Potamogeton vaginatus PI. A-3, fig. 6


Turcz. (sheathed
pondweed)

Fruits obliquely obovate, asymmetrical, 2.93.3 mm long, 2.2-2.6 mm wide, 1.4-1.8 mm thick.
Germination valve extends only two-thirds the length of the rounded dorsal margin and is not in contact
with the base of the slender style. Large seed strongly hook shaped and curved around condyle, seed
slightly smaller than locule. Seed coat smooth.

Zannichellia palustrisL. PI. A-3, fig. 7


var. palustris (horned:
pondweed)

Drupes oblong and curved, 2.0-2.1 mm long, 0.7-1.0 mm wide, 0.3-0.4 mm thick, with persistent pedicel
and style. Spine-like projections along margin on one side of drupes are unique to this species.

lncertae sedis (grass)

PI. A-3, fig. 8

Identificationof this grass material is restricted to the family level as only one vegetative fragment was
recovered and diagnostic organs are absent.

Carex cf. C. atherodes


Spreng. (awned sedge)

PI. A-4, fig. 1

Achenes narrow ovate in shape, trigonous, 2.1-2.7 mm long, 1.O-1.3 mm wide, 1.2-1.4 mm thick, sides
slightly concave. Maximum width is midway between centre of achene and base. Although a number of
important features are not preserved, comparison of fossil achenes with those of the living species
indicates that the fossils most closely resemble Carex atherodes.

Carex cf. C. rostrata


Stokes (beaked sedge)

PI. A-4, fig. 2

Achenes small, wide obovate, trigonous, 1.4-1.7 mm long, 1.O-1.3 mm wide, 0.8-1.3 mm thick. Base
acute with remnants of a stalk present, sides concave, margins rounded. Greatest width near the style.
Style attached to seed and angled, curled like a corkscrew, 1.8-1.9 mm long, 0.1-0.2 mm wide. Style
morphology and seed shape and size of the fossil are best compared to that of Carex rostrata achenes.

R. & S. (creeping spike-

Achene obovate, lenticular, 1.9 mm long, 1.0 mm wide, 0.5 mm thick. Seed broadest below tubercule and
narrowing toward base. Tubercule conical, longer than broad. Perianth composed of bristles of varying
length; bristles with retrorse barbs.

C.H. Yansa and J.F. Basinger

Specles

Flgure(s), Appendix A

Descrlptlon and remarks

Scirpus americanus
Pers. (three-square
bulrush)

PI. A-4, fig. 4

Achenes wide obovate, lenticular, 2.9-3.3 mm long, 2.0-2.3 mm wide, 1.O-1 .Imm thick. Seeds broadest
and thickest in upper third near style; style base short, apiculate; achene base sharply acute.

Scirpus cf. S. validus


Vahl. (great bulrush)

PI. A-4, fig. 5

Achenes wide obovate, lenticular, 1.7-2.1 mm long, 1.3-1.5 mm wide, 0.7-0.9 mm thick. Base of seed
broadly acute, base of style apiculate. Faint longitudinal furrows on outer surface of seed coat; epidermal
cells isodiametric, arranged in longitudinal rows, surface slightly textured. The identification of these fossil
seeds was difficult as the fossils are superficially similar in morphology to the seeds of Scirpus acutus
Muhl. Based on size, they are tentatively assigned to S. validus.

Lemna trlsulca L. (ivyleaved duckweed)

PI. A-4, fig. 6

Thallus elliptic to oblanceolate, flattened, 5.0-5.3 mm long, 1.5-2.1 mm wide; stipe 1.&3.1 mm long.
Venation faint, three veins visible; margin finely serrulate towards apex.

Wolffia arrhiza (L.)


Wimm. (water-meal)

PI. A-4, fig. 7

Thalli ovate to globose, free floating, rootless, 0.3-0.6 mm long, 0.3-0.5 mm wide. Distinctive border of
flattened, thick-walled epidermal cells.Young fronds of Wolffia arrhiza, produced from a basal cleft in
mature thalli, are also found preserved (PI. A-4, fig. 9).

Populus balsamifera L.
(balsam poplar)

PI. A-4, fig. 8 ( see also


Fig. 50)

Single terminal bud scale elongate, slender, up to 100 mm long, 2.4-2.7 mm wide, apex acute, margins
ciliate, surface smooth. Leaves ovate to ovate-lanceolate, tapering to an acute to acuminate to attenuate
apex, base broadly obtuse to cuneate to subcordate, 73.0-79.0 mm long, 56.0-60.0 mm wide. Venation
pinnate reticulodromous. Margins minutely crenulate to subentire.

Populus tremuloides
Michx. (aspen poplar)

PI. A-4, fig. 9 ( see also


Fig. 58)

Terminal bud scale slender, conical, 4.5-4.7 mm long, 2.4-2.8 mm wide, apex acumlnate, margins not
ciliated. Leaves broadly ovate to nearly circular, 33.0-53.0 mm long, 32.0-48.0 mm wide, apex acuminate,
base truncate to subcordate. Margins finely crenate to serrate to subentire.Venation reticulate
actinodromous. Petiole flattened and longer than blade.

Betula cf. B.
occidentalis Hook.
(river birch)

PI. A-5, fig. 1

Samaras elliptic to obovate, flattened, 1.I-1.3 mm long, 0.5-0.6 mm wide, lateral wings not preserved.
Styles two. Assignment to either Alnus or Betula is complicated by poor preservation.Tentatively assigned
to 6. occidentalis on the basis of size, because samaras of A. crispa are considerably larger than those of
6.occidentalis and the fossils.

Rumex marihmus L.
var. fueginus (Phil.)
Dusen (golden dock)

PI. A-5, fig. 2, 3

Achenes elliptic, symmetrical, sharply trigonous, 1.I-1.2 mm long, 0.5-0.6 mm wide, 0.4-0.6 mm thick,
with long marginal bristles; tubercles acute, 3.3-3.6 mm long, 2.1 mm wide, 2.0 mm thick. Closely
resembles those of R. maritimis var. fueginus seeds. Identification supported by recovery of associated
perianth segments with spiny margins (PI. A-5, fig. 5), considered characteristic of this species.

Chenopodium cf. C.
berlandieriMoquin
(false lamb's-quarters)

PI. A-5, fig. 4

Achene broadly obovate to nearly circular, asymmetrical, 1.O-1 .2 mm long, 0.9-1 .Imm wide, 0.4-0.7 mm
thick. On each side of the seed a shallow furrow extends radially from the centre to the hilum. Seed coat
has a reticulate pattern with occasional wart-like papillae. Morphology of the fossil seeds is superficially
similar to those of C. pratericola Rydb. and C. hians Standley. However, C, pratericola seeds are smaller
and more irregular in shape than the fossil seeds. Seeds of C.hians are circular and symmetrical,
whereas the micropylar end of the seed protrudes slightly in both the fossil and C. berlandieriseeds.
Fossil seeds therefore tentatively identified as C. berlandieri. Recently, C. berlandieri has been recognized
as a separate sp. and a native counterpart of C. album.

Chenopodium salinum
Standley (saline
goosefoot)

PI. A-5, fig. 5

Caltha palustris L.
(marsh-marigold)

PI. A-5, fig. 6

Achenes very wide obovate, asymmetrical, 0.7-0.8 mm long, 0.60.7 mm wide, 0.2-0.3 m m thick. Sides
slightly convex, seed margin distinct. An indistinct furrow extends one-third the length of the seed from the
hilum towards its centre. Seed coat texture slightly wavy to smooth. Seeds of C.rubrum L, var, rubrumare
similar in size and shape to the fossil seeds; however, they differ in being ovate in cross-section and
having a very smooth surface.
Seed significantly abraded, elliptic, nearly symmetrical. 2.7 mm long, 1.5 mm wide, 0.6 mm thick. Distinct
fold on one margin, apex truncate, margins entire, hilum swollen. Surface texture reticulate.

Ranunculus sceleratus
L. (celery-leaved
buttercup)

PI. A-5, fig. 7

Achenes very wide obovate, elliptic in cross-section, flat, nearly symmetrical, 0.9-1.0 mm long, 0.8-0.9
wide, 0.3-0.4 mm thick. Margin entire, slight furrow around periphery of seed visible, depression near
hilum, and outer keel obtuse with longitudinal groove. Surface texture faintly areolate. All other species of
Ranunculushave achenes that are significantly larger than the fossil and living R. sceleratus.

Rorippa cf. R. truncata


(Jepson) Stuckey
(blunt-fruitedyellow
cress)

PI. Ad. fig. 8

Seeds nearly circular, elliptic in cross-section,symmetrical, 0.8 mm long, 0.6-0.7 rnm wide, 0.3-0.4 mm
thick. Slight cleft between radicle and cotyledons near hilum. Surface texture reticulate to rugose. Seeds of
Rorippa palustris (Oeder) Borbus (= R. islandica (Oeder) Borbas) are closely comparable to fossil and
living R. truncata seeds and may easily be confused. However, R, truncata seeds are consistently nearly
circular in shape and possess rounded margins, whereas those of R, palustris are irregularly shaped and
have sharply ridged margins.

GSC Bulletin 534

Appendix B (cont.)
Species

Figure@), Appendix A

Description and remarks

Subularia aquatica L.
(water awl-wort)

PI. A-5, fig. 9

Seed narrow ovate, elliptic to oblong in cross-section, flattened, symmetrical, 0.8-1.0 mm long,
0.4-0.5 mm wide, 0.3 mm thick. Apex distinctly notched, broadest part of seed near middle. Longitudinal
primary groove between radicle and cotyledons centrally placed. Seed coat glabrous, slightly textured.

Fragaria virginiana
Dcne. (smooth wild
strawberry)

PI. A-6, fig. 1

Achenes ovate, obovate in cross-section, asymmetrical, 1.2-1.3 mm long, 0.8-1.0 mm wide, 0.5-0.7 mm
thick. Apex acute, slightly hooked adaxially. Faint striations radiating from hilum apparent on adaxial
surface. Surface texture smooth.This species of strawberry was identified at the Andrews site during the
floral survey.

Lemna sp. (duckweed)

PI. A-6, fig. 2

Achenes ovate to orbiculate, elliptic in cross-section, symmetrical, 0.7-0.8 mm long, 0.5-0.6 mm wide,
0.4-0.5 mm thick. Prominent indentation at one end and surface texture faintly areolate with prominent
longitudinal ribs are distinctive of this taxa.

Rubus idaeus L. (wild


red raspberry)

PI. A-6, fig. 3

Endocarps elliptic to suborbiculate, elliptic in long-section, asymmetrical, 1.8-2.4 mm long, 1.2-1.6 mm


wide, 0.9-1.0 mm thick. Features characteristic of this species include distinct suture line between two
halves, apex beak-like, and surface texture deeply alveolate with series of anastomosing ridges.

Geranium cf. G.
bicknellii Britt.
(Bicknell's geranium)

PI. A-6, fig. 4

Seeds narrow elliptic, circular in cross-section, symmetrical, 1.6 mm long, 0.5 mm wide, 0.5 mm thick.
Faint protuberances at both ends. Surface reticulate to rugose. Poor preservation of this material made
the identification of these fossil seeds difficult. However, the morphology of the fossil seeds most closely
resembles the seeds of G. bicknellii.

Shepherdia canadensis PI. A-6, fig. 5


(L.) Nun. (Canada
buffaloberry)

Achenes ovate, asymmetrical, wide elliptic in cross-section, 3.1-4.1 mm long, 2.0-3.0 mm wide,
0.8-0.9 mm thick. Prominently lobed on one side with longitudinal groove extending from base to apex.
Surface appearing smooth to slightly textured.

Epilobium cf. E.
ciliatum Raf. (northern
fireweed)

PI. A-6, fig. 6

Seeds narrow elliptic, transversely elliptic in cross-section, symmetrical, 1.0 mm long, 0.3 mm wide,
0.2 mm thick. Apex prominent. Based on surface texture, which appears longitudinally striate, and the
presence of an opening in the hilum, the fossil seeds are tentatively identified as those of E. ciliatum.

Hippuris vulgaris L.
(mare's-tail)

PI. A-6, fig. 7

Seeds oblong-cylindric to elliptic, circular in cross-section, symmetrical, 2.0-2.2 rnm long, 1.1-1.2 mm
wide, 1.O-1 . I rnm thick. Faint groove extending from base to apex. Large opening, up to 0.5 mm In
diameter at apex. Fruit wall thick and woody; surface texture areolate.

Myriophyllum sp.
(water-milfoil)

PI. A-6, fig. 8

Mericarps oblong to elliptic, oblong in cross-section, symmetrical, 2.1 mm long, 1.0 mm wide, 1.2 mm
thick. Apex truncated to slightly emarginate; base rounded. Margins slightly rounded and bearing small
tooth-like processes. Identification of fossil mericarps to the specific level is not possible.

Myriophyllum
verticillatum L. var.
pectinatum Wallr.
(whorled water-milfoil)

PI. A-6, fig. 9

Bract palmate, broadly obovate, 1.6-1.7 mm long, 1.I-1.3 mm wide, is diagnostic of only this species and
variety.

Lycopus americanus
PI. A-7, fig. 1
Muhl. (water-horehound)

Nutlets oblate to orbiculate, obliquely triangular In cross-section, symmetrical, 1.I-1.3 mm long,


0.8-0.9 mm wide, 0.4-0.5 mm thick. Central region of nutlet raised and surrounded by thick marginal
flange, which is characteristic of this species. Surface texture coarse.

Mentha arvensis L. var.


villosa (Benth.) S. R.
Stewart (field mint)

PI. A-7, fig. 2

Nutlets suborbiculate, elliptic in cross-section, symmetrical, 0.7-0.9 mm long, 0.5-0.6 mm wide,


0.4-0.5 mm thick. Apex rounded; base wedge shaped and ocellate. Surface texture faintly areolate. The
floral survey at the study area indicates that this species presently grows there.

Astercf. A. novaeangliae L. (New


England aster)

PI. A-7, fig. 3

Achenes narrow oblong, elliptic in cross-section, flattened, symmetrical, 1.3-1.4 mm long, 0.3-0.4 mm
wide. 0.1-0.2 mm thick. The longitudinally ribbed, densely pubescent fossil achenes most closely
resemble those of Asternovae-angliae. The only feature which is not well preserved in the fossil achenes
is the pappus rim.

Cirsium cf. C. muticum


Michx. (swamp thistle)

PI. A-7, fig. 4

Achenes oblong to elliptic, elliptic and compressed in cross-section, symmetrical, 2.2-2.5 rnm long,
0.6-0.7 mm wide, 0.5-0.6 mm thick. Prominent longitudinal ribs and fine longitudinal striations. Seed body
tapers to rounded base. Apex with prominent cup-shaped cartilaginous pappus rim surrounding remnant of
style. Although the pappus bristles are poorly preserved, other features of the achenes are diagnostic of
living C. muticum achenes.

The litliostratigraphic record of late


Pleistocene-Holocene environmental change at the
Andrews site near Moose Jaw, Saskatchewan
Alec E.

it ken', William M. ~ a s tand


~ , Abigail K. ~

u r t ~

Aitlcen, A.E., Last, W.M., and Burt, A.K., 1999: The lithostratigraphic record of late
Pleistocene-Holocene environmental change at the Andrews site near Moose Jaw,
Saskatchewan; in Holocene Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate Change on the Southern Canadian
Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological Survey of Canada, Bulletin 534,
p. 173-1 81.

Abstract: Four lithostratigraphic units are recognized in a 5.8 m dugout excavation within hummocky
terrain: diamicton (Ll), laminated organic-rich gypsite (L2), calcareous clayey silt (L3), and massive clayey
silt (L4). The basal diamicton is supraglacial till deposited within a shallow depression prior to 10 200 BP.
Subsequent melt-out of buried, stagnant glacial ice created a shallow lacustrine basin ca. 10 000 BP. A subsequent 1200-year interval of severe aridity (to ca. 8800 BP) is reflected by a dominance of chemical sedimentation in a saline to hypersaline basin. A bed of sandy mud enriched in charcoal flakes and charred seeds
at the top of this evaporitic unit records a period of drainage basin destabilization by prairie fires, which
resulted in increased runoff and erosion. Sediments reflect a gradual transition, between ca. 8000 and
6000 BP, from the high-salinity and chemical-dominated sedimentary environment to a lake characterized
by lower salinity and mainly fine-grained clastic sedimentation.

R6sum6 :

Quatre unit& lithostratigraphiques ont CtC reconnues dans une excavation de fosse-r6servoir
de 5,8 m en terrain bosselk : un diamicton (Ll); une gypsite larninte riche en matikres organiques (L2); un
silt argileux calcaire (L3);et un silt argileux massif (L4). Le diamicton de base est un till supraglaciaire qui a
ttC d6posC dans une dCpression peu profonde avant 10200 BP. La fonte ultkrieure de glace glaciaire enfouie
stagnante a engendrC un bassin lacustre peu profond a environ 10 000 BP. Un intervalle ult6rieur trbs sec
d'une durCe de 1 200 ans (jusqu'a environ 8 800 BP) se traduit par la prkdominance d'une sedimentation
chimique dans un bassin salin a hypersalin. Une couche de boue sablonneuse enrichie en flocons de charbon
de bois et en graines calcin6es au sommet de cette unit6 6vaporitique ttmoigne d'une pCriode de
destabilisation du bassin versant par des feux de prairie, qui se sont traduits par un ruissellement et une
Crosion accrus. Les stdiments reflbtent une transition graduelle entre environ 8 000 et 6 000 BP, B partir
d'un milieu ~Cdimentaireh salinitk 6levCe et A prkdominance chimique vers un lac caractCris6 par une
salinitC plus faible et une sedimentation clastique essentiellement 2 grains fins.

' Department of Geography, 9 Campus Drive, University of Saskatchewan, Saskatoon, Saskatchewan S7N 5A5

Department of Geological Sciences, University of Manitoba, Winnipeg, Manitoba R3T 2N2


Department of Earth Sciences, University of Waterloo, 200 University Avenue West, Waterloo, Ontario N2L 3G1

GSC Bulletin 534

INTRODUCTION

METHODS

Internally drained depressions within hummocky disintegration and ice-thrust moraines are common across much of the
northern Great Plains. These depressions can provide both
paleoecological data (Beaudoin, 1993,1996; Beaudoin et al.,
1996) and sedimentological and/or pedological data
(Vreeken, 1994, 1996) that record environmental change
during the late Pleistocene and Holocene. This paper examines environmental changes recorded by the physical and
chemical sedimentology of one such site near Moose Jaw in
southern Saskatchewan (Fig. I), and complements the examination of plant macrofossils from this same site (Yansa,
1995, 1998; Yansa and Basinger, 1999).

A 5.8 m thick section was exposed at this site during the excavation of a dugout in 1993.

PHYSICAL SETTING
The Andrews site is located on The Missouri Coteau upland
in southern Saskatchewan (lat. 5020'N, long. 105"52'W,
elev. 720 m a.s.l.), 22 krn southwest of Moose Jaw (Fig. 1).
The sediments occupy an internally drained depression in
hummocky disintegration moraine. The regional climate is
strongly continental. Mean annual temperature is 3.gC, with
mean January and July temperatures of -14.2"C and 19.7OC,
respectively, and mean annual precipitation is 357 mm
(Longley, 1972; Environment Canada, 1993). Potential
evapotranspiration often exceeds precipitation by more than
300 rnrn annually (Winter, 1989).

Seventy-five sediment samples provided by C.H. Yansa


were analyzed for moisture content, organic matter content,
total carbonate content, particle size, and mineralogy. A regular sampling interval of approximately 10 cm was used,
except where bedding or other sedimentary features dictated
closer spacing. Moisture content, organic matter content, and
total carbonate mineral content were determined by loss on
ignition at 85"C, 500C, and 1000C, respectively (Dean,
1974). Particle-size spectra for each subsample were determined using an automatic laser-optical particle-size analyzer
(GALA1 CIS-lTM;Aharonson et al. (1986), Syvitski et al.
(1991)) after removal of organic matter by hydrogen peroxide
treatment. A 1% solution of TritonTMlOOX was used as the
dispersing agent, rather than disodium hexametaphosphate,
because of the high sulphate mineral content of many of the
samples. Particle-size statistics (mean, median, and standard
deviation) were calculated by the moment method according
to Allen (1981). Duplicate samples were prepared and analyzed every 20 samples, and a standard was run every 50 samples. These replicate analyses indicate that precision of the
particle-size data is approximately k 4%.
Samples analyzed for mineralogy were air dried at room
temperature, disaggregated in a mortar and pestle, and passed
through a 62.5 Fm sieve. Bulk mineralogy and detailed

t o - + End moraine

----

-LY

Ice-thrust ridges
Meltwater channel

Figure 1. Location of the study site (indicated by a star) in hummocky moraine in southern
Saskatchewan (modified from Saskatchewan Research Council, 1987).Surficial deposits of the study
area: Ap, alluvial plain; GLp, GLv, G M , GLh, glaciolacustrine (plain, veneer, drumlinoid, and
hummocky respectively); GFp, GFh, glaciojluvial (plain and hummocky, respectively); Mh,
hummocky moraine; Mr, ice-thrust ridge; Mb, hummoclcy disintegration moraine; Mp, Mu, Me, Mg,
ground moraine (plain, undulating, eroded, and gullied, respectively); Eh = eolian (hummocky).

A.E. Aitken et al.

600
0

2000

4000

6000

8000

10000

I4C age (years BP)

Figure 2. Depth versus radiocarbon age in the section from


the Andrews dugout.

Moisture

Organic
matter

Texture

carbonate and evaporite mineralogy were determined using


standard X-ray diffraction (XRD) techniques (Miiller, 1967;
Klug and Alexander, 1974). Mineral identification was aided
by the use of an automated search-match computer program
(Marquart, 1986). Percentages of the various minerals were
estimated from the bulk mineral diffractograms, using the
intensity of the strongest peak for each mineral as outlined by
Schultz (1964). Nonstoichiometry of the dolomite and calcite
was determined by measuring the displacement of the d,,,
peak on a detailed (slow) XRD scan (Goldsmith and Graf,
1958) and calculated according to Hardy and Tucker (1988).
and analyzed for bulk minDuplicate samples were
eralogy every 20 samples. These replicate analyses indicate
that the precision of the mineralogical data is approximately
k 8%. The abundances of the feldspar minerals were used to
calculate an index of weathering. This index responds to the
chemical stabilities of potassium feldspars (high stability)
versus plagioclase feldspars (low stability) in a typical drainage-basin weathering profile. Although many factors influence the abundance of these detrital siliciclastic minerals, the
ratio of K-feldspar to plagioclase in the offshore sediments of
a small lake with a relatively homogeneous drainage basin
should increase under more intense chemical weathering (i.e.
warm temperatures, humid climate) and decrease under less
severe conditions (i.e. cooler, drier climate).

Mean
size

Standard
deviation of
mean size

Median
size

I'm

I'm

,;

,-:
;,7T,;Missing samples
50
%

Yo

1001

10

I'm

100

Figure 3. Stratigraphic variation in moisture content, organic matter content, texture (%sand, %silt, % clay),
and particle size (mean, standard deviation, and median). Letters and numbers in capitals on the right side of
the figure (e.g. LI) and horizontal dashed lines refer to lithostratigraphic units.

GSC Bulletin 534

Both conventional and accelerator mass spectrometry


(AMS) radiocarbon ages were acquired for this study by
Yansa (1995, 1998) (see Yansa and Basinger, 1999). The
dearth of suitable organic materials in the upper 300 cm of the
measured section precluded dating of these sediments
(Yansa, 1998). All dates are reported here as radiocarbon
years before present. The chronostratigraphy employed in
this paper was established by fitting a third-order polynomial
to the 14Cdates and assuming continuous, uninterrupted sedimentation to present and no postdepositional compaction
(Fig. 2).

trend in mean grain size (Fig. 3; Table 1). The lowermost 2 m


of sediment are somewhat coarser grained and more poorly
sorted (higher standard deviation) than the rest of the section,
while the upper 2 m of the section exhibit regular alternation
between relatively fine-grained and coarse-grained sediments (Fig. 3). Moisture and organic matter contents are both
relatively low (<30% and <lo%, respectively), except
between 4 and 5 m depth.
The sediments are composed predominantly of detrital
clay minerals (46%) and detrital siliciclastic minerals (30%),
with smaller proportions of detrital carbonate minerals (8%)
and endogenic and authigenic carbonates and evaporites
(16%) (Table 1). Detrital sediments dominate except for the
interval between about 4 and 5 m depth (Fig. 4), which contains a higher proportion of endogenic and authigenic components. The endogenic and authigenic fractions of the
sediment consist mainly of gypsum (CaS04.2H,0), with
minor amounts of disordered, nonstoichiometric dolomite
(Ca0,xMg0.y(C03)2)
and magnesian calcite (Table 1, Fig. 5).
Other carbonate and sulphate salts occur sporadically within

RESULTS
General description
The exposed section consists of 80 cm of diamicton overlain
by 500 cm of nonglacial sediments. Overall, there is little systematic change in sediment texture (average composition: 4%
sand, 67% silt, 29% clay), other than a slight fining-upward

Table 1. Summary of average mineralogical and grain-size characteristics


of lithostratigraphic units.
Lithostratlgraphlc unlt

I %total clay minerals

L1

128.7

L2
38.9

L3

52

L4

Core
average

56

45.9

<0.1

c0.1

14

19.3

% quartz
% totai feldspar mlnerals

% plagioclasefeldspars
% K-feldspars
% amphibole minerals
% total carbonate mlnerals
%total detrital carbonate minerals
% total endogenic + authlgenic carbonate minerals
% total calcite
% low-Mg calcite
% magnesian calcite
Mol % MgC03in magnesian calcite
% totai dolomite
% well-ordereddolomite
% dlsordered dolomite
Mol % CaC03 In protodolomite
( % magnesite
% aragonite
%gypsum
% (Mg + Na)-sulphate salts
% Na-sulphatesalts
% detrital fraction
D/. endogenic + authigenlc fraction
% moisture
YOorganic matter
Mean particle size (pm)
Standard deviation of mean particle size
Median particle size (pm)
% sand
% silt
%clay
Sedimentation rate1(cm.(100 a).')

n.p.

n.p.

n.p.

32.1

22.6

10.3

9.4

5.5

5.1

7.3

9.6

5.4

2.1

4.2

66.7

73.8

66

62

67.1

23.7

20.8

31

36

28.8

n.d.

4.3

18

3.4

5.1

4.6

n.d. = not determined; n.p. = not present.


Calculated on the basis of a best-fit line of depth versus time (Fig. 1) using the six 14cdates of Yansa
11997) and assumina no interruotion in deoosition.

'

A.E. Aitken et al.

the section; these include bloedite (Na,Mg(SOd)3.4H,0),


thenardite (Na2S04), or the hydrated species 'mirabTlite
(Na2S04.10H20), aragonite (CaC03), and magnesite
(MgC03) (Fig. 5).
Over the entire section, variations in the proportions of the
various siliciclastic components (quartz, K-feldspar,
plagioclase, amphiboles) are all significantly correlated with
one another (linear correlation coefficients are significant at
the 95% confidence level) and with the various-grain-size
parameters, and inversely correlated with the abundance of
organic matter, gypsum, and the detrital carbonates. Similarly, detrital calcite and dolomite are covariant and are negatively correlated with the endogenic carbonate minerals. The
abundance of gypsum has a significant positive correlation
with mean and median grain size and a negative correlation
with clay content, suggesting that this soluble evaporite mineral is dominant in the coarser fraction of the sediment.

L1- Diamicton
Extending from the base of the section to 500 cm depth, this
unit consists of slightly gravelly, sandy mud with low moisture and organic matter contents. In addition to the relatively
coarse mean grain size and high sand content, this poorly
sorted detrital unit is marked by high detrital carbonate
content and the absence of endogenic and authigenic
components.

A coarse organic debris layer (510-500 cm depth), indicated by a sharp increase in sediment organic matter content
at the top of L1 (Fig. 3), consists largely of macrofossils of
Picea glauca (white spruce), Rubus idaeus (raspberry),
Shepherdia canadensis (soapberry), and charcoal flakes,
charred wood, and seeds in a sandy mud matrix (Yansa,
1995). Roots associated with white spruce trunks penetrate
these organic-rich sediments. Both the roots and trunks have
been radiocarbon dated at 10 200 BP (Yansa, 1995, 1998;
Yansa and Basinger, 1999).

Lithostratigraphy
The four lithostratigraphic units recognized in the section
(Table 1, Fig. 3-5)-are described bel;w (in order from the
base of the section).

L2 - Organic-rich gypsite
A 1 m thick (400-500 cm depth), thinly laminated (laminae
2-5 mm thick), organic-rich evaporite unit overlies the
diarnicton. The clayey silt is characterized by high gypsum,

Miss~ngsamples

-(Diamicton)
Ll

Missing samples

w-

Figure 4. Stratigraphic variation in detrital quartz,feldspar minerals (potassiumfeldspars, plagioclase


feldspars, total feldspars), total clay minerals, total carbonate minerals, and the proportion of detrital
versus (endogenic + authigenic) mineral components. Letters and numbers in capitals on the right side
of the figure (e.g. L1) and horizontal dashed lines refer to lithostratigraphic units.

GSC Bulletin 534

minerals differs from that in the underlying unit: the mol %


MgC03 within the magnesian calcite is relatively high and
the mol % CaC03 content of the protodolomite is low.
Although gypsum is rare, unit L3 does contain sporadic
occurrences of other sulphate salts. Sediment accumulation
in L3 occurred during the interval 8150-5500 BP.

magnesian calcite, and protodolomite contents, together with


detrital calcite and dolomite. Discrete thin beds and laminae
of fine to medium quartzose sand occur over the interval
470420 cm within this unit.
Sediment accumulation in L2 occurred during the interval
ca. 10 200-8150 BP. The laminated sediments are overlain by
a 10 cm thick bed (390400 cm depth) of sandy mud enriched
in charcoal flakes and charred seeds. This horizon is associated with a pronounced reduction in the organic matter content, from about 20% to less than 5%.

L4 - Clayey silt
The upper approximately 3 m of the section consists of
nonbedded, noncalcareous, clayey silt to silty clay. High
quartz, feldspar, and clay mineral contents, low organic matter content, and an almost complete absence of detrital carbonates characterize the composition of this unit. Slightly
coarser grained sediments occur in the intervals 230-295 cm
and 30-150 cm. Variation in mean grain size and standard
deviation in the upper 3 m suggest a regular alternation of relatively fine and relatively coarse sediment, although this is
not reflected by any visible bedding features.

L3 - Calcareous clayey silt


This unit, extending from 390 to 295 cm depth, is a
nonbedded, calcareous, clayey silt. The unit is distinguished
from underlying L2 by its finer texture, high clay-mineral
content, lack of bedding, and paucity of gypsum. As in L2,
magnesian calcite and protodolomite dominate the endogenic
carbonates. However, the composition of these carbonate

..
.
.
. . ...
..
c.
..
....

0 ,,,,,,,,,,

100

200

, , , , , , , , , I

I I I I I I I I
A

*.

Mg-Calcite

300

------L--

P
Q)

n
400

------

r.

d
500

tcalcite

-Na-sulphates

.
--A

600

-----r A -

l l l l l l l l a l

10

20

T;,7y;y
10
Mol %
MgCO,

20

10

YO

20

65
Mol %
CaCO,

80

40

80

20

40

Weathering
Intensity

Figure 5. Stratigraphic variation in calcite, magnesian calcite, dolomite, protodolomite, gypsum,


Na-sulphates, and (Na+Mg)-sulphates.Also shown are variations in the composition of the magnesian
calcite and protodolomite, and the clastic weathering index. The weathering index is a ratio of the relatively chemically stable potassium feldspars to the less stable plagioclase feldspars. This ratio increases
(to the right in the figure) under more intense chemical-weathering conditions in the watershed. Letters
and numbers in capitals on the right side of the figure (e.g. L1) and horizontal dashed lines refer to
lithostratigraphic units .

A.E. Aitken et al.


-

INTERPRETATION
The basal diarnicton (Ll) is most likely Late Wisconsinan
Battleford Formation, on the basis of stratigraphic position
and lithological characteristics (cf. Christiansen, 1971,1992;
Klassen, 1992). Subglacial erosion of Cretaceous marine
sandstone and shale and Paleozoic carbonate rocks beneath
the Laurentide Ice Sheet is reflected in the dominance of
detrital carbonate minerals and siliciclastics in this
lithostratigraphic unit. Like elsewhere on The Missouri
Coteau, thick supraglacial debris resulting from thrusting at
the ice margin would have insulated the glacier surface from
direct radiation and may have allowed stagnant ice to persist
for several thousand years following retreat of active ice from
the site. Down-wasting of the ice surface, accompanied by
active resedimentation of supraglacial debris, generated a
multitude of internally drained depressions within
hummocky moraine (Gravenor and Kupsch, 1959; Clayton,
1967).
Coarse organic debris present at the top of L l (vegetation
zone I1 of Yansa and Basinger, 1999), together with charcoal
and charred wood and seeds, suggest fire may have temporarily disrupted the vegetation cover of the landscape surrounding the depression ca. 10 200 BP. The resulting
destabilization of watershed soils would serve to increase
runoff and erosion.
The sharp contact between the poorly sorted mineral and
organic detritus of L1 and the fine-grained laminated
evaporites of L2 implies a relatively rapid transition to
high-salinity lacustrine conditions. The fine, undisturbed
lamination, relatively fine grain size, and high organic-matter
content of L2 also indicate deposition in quiet water conditions. Although these sedimentary features are similar to
those characterizingpresent-day offshore (deep-water) facies
of perennial and meromictic saline lakes in the region
(Schweyen and Last, 1983; Van Stempvoort et al., 1993;
Valero-GarcCs et al., 1995), finely laminated, organic-rich
evaporitic sediments can also be generated and preserved in
shallow playa basins (Last and Vance, 1997).
The endogenic mineral suite of L2 is dominated by gypsum, protodolomite, and magnesian calcite. Comparison with
modern analogues on the northern Great Plains indicates that
the genesis of these minerals occurs in basins having salinities ranging from about 10 to 40 g - ~ -(Last,
l
1994; Last and
Vance, 1997;Vance et al., 1997). Variation in ionic composition of the lake water, from Ca-sulphate dominated to
(Mg+Ca)-bicarbonate enriched, is required to facilitate the
contemporaneous deposition of gypsum and carbonates.
High MgICa ratios (10-100) and elevated alkalinities are
indicated by the presence of magnesian calcite and
protodolomite. Sulphate salts also occur sporadically within
the uppermost 20 cm of L2 and occasionally in overlying L3,
suggesting that the basin experienced brief periods of
hypersalinity (TDS greater than 40 g - ~ - l in
) the early
Holocene. Although there is nothing in the sediment to indicate complete drying and desiccation, the lake likely experienced significant fluctuations in water level, from deep and
possibly meromictic conditions to shallow ephemeral

conditions, during this 1200-year period. This interval was


terminated abruptly by deposition of a 10 cm thick bed
(390-400 cm depth) of sandy mud enriched in charcoal flakes
and charred seeds. This episode, which corresponds to vegetation zone IV of Yansa and Basinger (1999), is also interpreted to reflect an increase in erosion due to prairie fires
destabilizing the vegetation cover of the watershed. This
period of increased detrital accumulation is bracketed by
radiocarbon ages of 8800 BP at 410 cm depth and 7700 BP at
395 cm depth.
The abundance of endogenic minerals in L3 is lower than
that of L2, but the presence of both Mg-calcite and
protodolomite in L3 indicates continued high alkalinity and
moderate salinity. The significant increase in mol % MgC03
in the endogenic calcite (and a corresponding decrease in calcium content in protodolomite) of L3 relative to L2 suggests
an increase in MgJCa ratios. The gradual upward increase in
allogenic components continues into the upper 3 m of the section (unit L4). This upper portion is also characterized by an
increase in the weathering intensity index, suggesting more
humid and/or warmer climatic conditions after ca. 5900 BP.
Overall, however, the mid- to late Holocene record in this
sequence does not appear to be especially sensitive to environmental changes. Except for broadly spaced shifts in mean
grain size and sorting, and the rare occurrence of sulphates,
the mineralogical composition and organic matter content of
L4 are homogeneous.

DISCUSSION
The mineralogical and textural data from the Andrews site
generally correspond well with the interpreted lake history
deduced from paleobotanical evidence at the site (Yansa,
1995,1998; Yansa and Basinger, 1999). Nonetheless, differences between the two records are apparent. Yansa (1998)
and Yansa and Basinger (1999) proposed that the lake shifted
gradually from a permanent, deep-water (>2 m) pond environment ca. 10 000 BP to an alkaline and brackish, shallow,
semipermanent wetland by 8800 BP. In their reconstruction,
the local onset of maximum Holocene aridity did not occur
until after 8800 BP. This interpretation is not supported by the
mineralogical data presented here. Rather, the rapid transition
to highly saline conditions at the L2-L3 boundary (ca.
10 000 BP) suggests an earlier phase of severe aridity that is
consistent with records of several other sites in the region. For
example, lake records from three sites within the Palliser
Triangle all suggest an interval of severe aridity in the Late
Pleistocene-early Holocene. At Clearwater Lake, about
100 km northwest of the Andrews site, Last et al. (1998)
recognized a saline to hypersaline phase (salinities approximately 20-70 g . ~ - lbetween
)
9800 and 8600 BP. Oro Lake,
located about 50 krn south of the Andrews site, recorded an
abrupt change from fresh-water conditions to a hypersaline,
meromictic lake ca. 9400-9000 BP (Vance and Last, 1996).
Salts preserved in North Ingebrigt lake, a hypersaline playa
about 250 km west of the Andrews site, document much
lower atmospheric relative humidities ca. 9800 BP (Shang
and Last, 1999). Elsewhere on the Canadian Prairies, a shift
from fresh to saline conditions occurred between 10 000 and

GSC Bulletin 534

9000 BP at Moore Lake, in east-central Alberta (Hickman


and Schweger, 1993). In the United States, a salinity maximum was recorded between 10 200 and 9500 BP in Pickerel
Lake, South Dakota, one of the longest stratigraphic records
from the Great Plains (Schwalb and Dean, 1998). Similarly,
Radle et al. (1989), Kennedy (1994), and Valero-GarcCs et al.
(1995) showed that Medicine Lake, South Dakota was a
highly saline basin at 9500 BP, while Xia et al. (1997) interpreted an increase in salinity ca. 9000 BP in Coldwater Lake,
North Dakota.
Such differences, however subtle, illustrate the value of
interpretations based upon integration of multiple-proxy
indicators. Although neither changes in sediment mineralogy
nor plant macrofossils are a direct record of climate change,
the mineralogical record does allow detailed reconstruction
of past brine chemistry. And although brine chemistry may
relate, in part, to changes in groundwater regime (e.g.
Remenda and Birks, 1999), the setting of the Andrews site
suggests that climate will remain a dominant factor. It is noteworthy that the peak salinities observed in the mineralogical
record at the Andrews site correlate well with the peak occurrences of Chenopodium salinum (saline goosefoot; cf. Fig. 4
in Yansa and Basinger (1999)), one of the most important
halophytic species on the Prairies. The fact that signals of
maximum aridity from different proxy indicators may occur
at different times at the same site largely reflects the differing
response times of hydrological, ecological, and geomorphic
systems.

CONCLUSIONS
The transition from cool, moist, late-glacial environments to
warmer and drier, early Holocene environments is recorded
in the lithostratigraphy of postglacial sediments at the
Andrews site. Melt-out of stagnant glacial ice and the collapse of overlying supraglacial debris contributed to the formation and initial flooding of a lacustrine basin in hummocky
moraine prior to 10 200 BP. Laminated, fine-grained lacustrine sediments, rich in organic matter and endogenic precipitates, were deposited in this basin over the interval ca.
10 000-8000 BP. Salinities may have reached 40 g . ~ - l ,
alkalinities were high, and the water was dominated by M ~ ~ + ,
ca2+, and ~
0 ions.~ After
~ ca. -8000 BP, the water composition changed to generally lower salinity, with little sulphate
but high Mg/Ca ratios. The increasing relative abundance of
detrital clay minerals, quartz, and feldspar minerals in these
sediments reflects an increase in the rate of erosion in the
lake's watershed and corresponds to an increase in the intensity of chemical weathering, probably due to a warmer and
more humid climate after ca. 8000 BP.

ACKNOWLEDGMENTS
The authors wish to acknowledge the financial support provided by the Geological Survey of Canada (GSC), the Natural
Science and Engineering Research Council of Canada
(NSERC), the University of Saskatchewan, and the

University of Manitoba. Catherine Yansa, Department of


Geography, University of Wisconsin-Madison provided the
samples examined in this study. Pat Athanasopoulos, Grant
Carelse, and Anne Ross provided valuable assistance in the
laboratory. Cartographic production was undertaken, in part,
by Keith Bigelow and Nathalie Stevenson, Department of
Geography, University of Saskatchewan. The manuscript has
benefited from reviews provided by Alwynne Beaudoin,
Donald Lernmen, Robert Vance, and Rod McGinn.

REFERENCES
Aharonson, E.F., Karasikov, N., Roitberg, M., and Shamir, J.
1986: GALAI-CIS-1. A novel approach to aerosol particle size analysis;
Journal of Aerosol Science, v. 17, p. 530-536.
Allen, T.
1981: Particle Size Measurement; Chapman & Hall, Toronto, Ontario,
468 p. (third edition).
Beaudoin, A.B.
1993: A compendium and evaluation of postglacial pollen records in
Alberta; Canadian Journal of Archaeology, v. 17, p. 92-112.
1996: An early Holocene macrofossil record from the Fletcher site
(DjOw-1) in southern Alberta; Canadian Association of Geographers Annual Meeting, Program and Abstracts, p. 6 1 4 2 .
Beaudoin, A.B., Yansa, C.H., and Vance, R.E.
1996: A model for late glacial-early Holocene landscape developmenton
the northern Great Plains; Canadian Association of Geographers
Annual Meeting Program and Abstracts, p. 62-63.
Christiansen, E.L.
1971: Tills in southern Saskatchewan; in Till: A Symposium, (ed.)
R.P. Goldthwaite; Ohio State University Press, Columbus, Ohio,
p. 167-183.
1992: Pleistocene stratigraphy of the Saskatoon area, Saskatchewan,
Canada: an update; Canadian Journal of Earth Sciences, v. 29,
p. 1767-1778.
Clayton, L.
1967: Stagnant-glacierfeatures of The Missouri Coteau in North Dakota;
in Glacial Geology of The Missoui Coteau and Adjacent Areas,
(ed.) L. Clayton and T.F. Freers; North Dakota Geological Survey,
Miscellaneous Series 30, p. 2 5 4 6 .
Dean, W.E.
1974: Determination of carbonate and organic matter in calcareous sediments and sedimentary rocks by loss on ignition: comparison with
other methods; Journal of Sedimentary Petrology, v. 44,
p. 242-248.
Environment Canada
1993: Climatic normals, 1961-1990: Prairie Provinces; Supply and Services Canada, Ottawa, Ontario, 266 p.
Goldsmith, J.R. and Graf, D.L.
1958: Relation between lattice constants and co~npositionof the Ca-Mg
carbonates; American Mineralogist, v. 43, p. 84-101.
Gravenor, C.P. and Kupsch, W.O.
1959: Ice-disintegrationfeatures in western Canada; Journal of Geology,
V. 67, p. 48-64.
Hardy, R. and Tucker, M.
1988: X-ray powder diffraction of sediments; in Techniques in
Sedimentology,(ed.) M. Tucker; Blackwell Scientific Publications,
Boston, Massachusetts, p. 191-228.
Hickman, M. and Schweger, C.E.
1993: Late glacial-early Holocene palaeosalinity in Alberta, Canada climate implications;Journal of Paleolimnology,v. 8, p. 149-160.
Kennedy, K.A.
1994: Early Holocene geochemical evolution of saline Medicine Lake,
South Dakota; Journal of Paleolimnology, v. 10, p. 69-84.
Klassen, R.W.
1992: Nature, origin, and age relationships of landscape complexes in
southwestern Saskatchewan; GCographie physique et Quaternaire,
v. 46, p. 361-388.
Mug, H.P. and Alexander, L.E.
1974: X-ray Diffraction Procedures for Polycrystalline and Amorphous
Materials; John Wiley and Sons, New York, 456 p.

A.E. Aitken et al.


Last, W.M.
1994: Deep-water evaporite mineral formation in lakes of western
Canada; in Sedimentology and Geochemistry of Modern and
Ancient Saline Lakes, (ed.) R. Renaut and W.M. Last; Society of
Economic Paleontologists and Mineralogists, Special Publication
NO.50, p. 5 1-60.
Last, W.M. and Vance, R.E.
1997: Bedding characteristics of Holocene sediments from salt lakes of
the northern Great Plains, western Canada; Journal of
Paleolimnology, v. 17, p. 297-318.
Last, W.M., Vance, R.E., Wilson, S., and Smol, J.P.
1998: A multi-proxy limnologic record of rapid Early Holocene
hydrologic change on the northern Great Plains, southwestern
Saskatchewan, Canada; The Holocene, v. 8, p. 503-520.
Longley, R.W.
1972: The climate of the Prairie Provinces; Environment Canada, AtrnosphericEnvironment Services,Climatological StudiesNo. 13,79p.
Marquart, R.G.
1986: pPDSM: mainframe searchlmatch on an IBM PC; Powder Diffraction, v. 1, p. 34-36.
Miiller, G.
1967: Methods in Sedimentary Petrology; Part I of Sedimentary Petrology, W.V.Engelhardt, H. Fiichtbauer and G. Miiller; Hafner Publishing Company, New YorklLondon, 276 p. (translated by
H-U. Schminke).
Radle, N., Keister, C.M., and Battarbee, R.W.
1989: Diatom, pollen, and geochemical evidence for the paleosalinity of
MedicineLake, South Dakota, during theLate Wisconsin and early
Holocene; Journal of Paleolimnology, v. 2, p. 159-172.
Remenda, V.H. and Birks, S.J.
1999: Groundwater in the Palliser Triangle: an overview of its vulnerability and potential to archive climate information; in Holocene Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Saskatchewan Research Council
1987: Surficial geology of the Regina area (72I), Saskatchewan;
Saskatchewan Research Council, scale 1:250 000.
Schnltz, L.G.
1964: Quantitative interpretation of mineralogical composition from
X-ray diffraction and chemical data for the Pierre Shale; United
States Geological Survey, Professional Paper 391-C, 31 p.
Schwalb, A. and Dean, W.E.
1998: Stable isotopes and sediments from Pickerel Lake, South Dakota,
USA: a 12 ka record of environmental changes; Journal of
Paleolimnology, v. 20, p. 15-30.
Schweyen, T.H. and Last, W.M.
1983: Sedimentology and paleohydrology of Waldsea Lake,
Saskatchewan; Canadian Plains Research Centre, University of
Regina, Regina, Saskatchewan,Canadian Plains Proceedings,v. 11,
p. 45-59.
Shang, Y. and Last, W.M.
1999: Mineralogy, lithostratigraphy, and inferred geochemical history of
North Ingebrigt lake, Saskatchewan; in Holocene Climate and
Environmental Change in the Palliser Triangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.

Syvitski, J.P.M., LeBlanc, K.W.G., and Asprey, K.W.


1991: Interlaboratory, interinstrument calibration experiment; in Principles, Methods, and Applications of Particle Size Analysis, (ed.)
J.P.M. Syvitski; Cambridge University Press, Cambridge, United
Kingdom, p. 174-193.
Valero-Garch, B.L., Kelts, K., and Ito, E.
1995: Oxygen and carbon isotope trends and sedimentological evolution
of a meromictic and saline lacustrine system: the Holocene Medicine Lake basin, North American Great Plains, U.S.A.; Palaeogeography, Palaeoclimatology,Palaeoecology, v. 117, p. 253-278.
Vance, R.E. and Last, W.M.
1996: Stop 4: Oro Lake; in Landscapes of the Palliser Triangle, Guidebook for the Canadian Geornorphology Research Group Field Trip,
(ed.) D.S. Lemmen; Canadian Association of Geographers, 1996
Annual Meeting, Saskatoon, Saskatchewan, p. 26-27.
Vance, R.E., Last, W.M., and Smith, A.J.
1997: Hydrologicand climatic implications of a multidisciplinarystudy of
late Holocene sediment from Kenosee Lake, southeastern
Saskatchewan, Canada; Journal of Paleolimnology, v. 18,
p. 365-393.
Van Stempvoort, D.R., Edwards, T.W.D., Evans, M.S., and Last,
W.M.
1993: Paleohydrology and paleoclimate records in a saline prairie lake
core: mineral, isotope, and organic indicators; Journal of
Paleolimnology,v. 8, p. 135-147.
Vreeken, W.J.
1994: A Holocene soil-geomorphic record from the Ham site near Frontier, southwestern Saskatchewan; Canadian Journal of Earth Sciences, v. 31, p. 532-543.
1996: A chronogram for postglacial soil-landscape change from the
Palliser Triangle, Canada; The Holocene, v. 6, p. 433438.
Winter, T.C.
1989: Hydrologic studies of wetlands in the northern Prairies; in Northern
Prairie Wetlands, (ed.) A. van der Vak; Iowa University Press,
Ames, Iowa, p. 17-54.
Xia, J. Haskell, B.J., Engstrom, D.R., and Ito, E.
1997: Holocene climate reconstructions from tandem trace-element and
stable-isotope composition of ostracodes from Coldwater Lake,
North Dakota, U.S.A.;Journal of Paleolimnology,v. 17, p. 85-100.
Yansa, C.H.
1995: An early postglacial record of vegetation change in southern
Saskatchewan, Canada; M.Sc. thesis, University of Saskatchewan,
Saskatoon, Saskatchewan, 273 p.
1998: Holocene paleovegetation and paleohydrology of a prairie pothole
in southern Saskatchewan, Canada; Journal of Paleolimnology,
v. 20, p. 429441.
Yansa, C.H. and Basinger, J.
1999: A postglacial plant macrofossil record of vegetation and climate
change in southern Saskatchewan; in Holocene Climate and Environmental Change in the Palliser Triangle: A GeoscientificContext
for Evaluating the Impacts of Climate Change on the Southern
Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological
Survey of Canada, Bulletin 534.

Sand dunes of the northern Great Plains of Canada


and the United States
Daniel R. ~ u h s and
' Stephen A. wolfe2
Muhs, D.R. and Wove, S.A., 1999: Sand dunes of the northern Great Plains of Canada and the
United States; in Holocene Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Contextfor Evaluating the Impacts of Climate Change on the Southern Canadian
Prairies, (ed.) D.S. Lemrnen and R.E. Vance; Geological Survey of Canada, Bulletin 534,
p. 183-197.

Abstract: Mostly stable dune fields are widespread over the subhumid to semiarid northern Great Plains.
Although winds in the region are strong, most dunes are presently inactive because of relatively high ratios
of precipitation to potential evapotranspiration, which has the dual effect of increasing dune moisture content and maintaining a vegetation cover. Dune fields are relatively small, as most are derived from finite supplies of glaciofluvial or glaciolacustrine sediments from the last deglaciation. Many dunes, however, are not
relict features from the last deglaciation. The last episodes of eolian activity were during the late Holocene,
although there is as yet little evidence for regional synchroneity of sand movement. The strong winds and
negative moisture regime have combined to produce an eolian system that is highly sensitive to small shifts
in climate. The potential for reactivation of northern Great Plains sand dunes is great, whether due to natural
climatic variations or human-induced greenhouse warming.
RCsumC : De nombreux champs de dunes, dont la plupart sont stables, parsbment les grandes plaines
septentrionales subhumides 8 semi-aides. Malgre la force des vents dans la region, la plupart des dunes
sont actuellement inactives en raison du rapport relativement ClevC des prtcipitations 8 1'Cvapotranspiration
potentielle, ce qui a pour double effet d'accroltre le contenu en humidit6 des dunes et de maintenir une
couverture v6gCtale. Les champs de dunes sont relativement petits, car la plupart proviennent d'un
approvisionnement fini de s6diments fluvioglaciaires ou glaciolacush-es de la dernibre deglaciation.
Toutefois, de nombreuses dunes ne sont pas des vestiges de la dernikre deglaciation. Les derniers episodes
d'activit6 Colienne ont en effet eu lieu au cours de 1'Holocbne supkrieur, bien qu'on ne dispose pas jusqu'h
present d'indices de synchronisme regional des dkplacementsdu sable. L'action combinde des forts vents et
du r6gime d'humidit6 n6gative ont engendrt un systbme Colien qui est trbs sensible 8 de ICgbres modifications du climat. I1 existe un fort potentiel de reactivation des dunes de sable des grandes plaines
septentrionales, que ce soit en raison de variations climatiques naturelles ou d'un rechauffement par effet de
serre d'origine anthropique.

' United States Geological Survey, M.S. 980, Box 25046, Federal Center, Denver, Colorado, U.S.A. 80225
Terrain Sciences Division, Geological Survey of Canada, 601 Booth Street, Ottawa, Ontario KIA 0E8

GSC Bulletin 534

INTRODUCTION
Sand dunes, dominated by parabolic forms, and associated
eolian sand sheets are common landforms on the northern
Great Plains of Canada and the United States. The dunes and
eolian sand sheets of both southern Canada and the United
States were first mapped on a continental scale by Thorp and
Smith (1952). David (1977) conducted the first detailed studies of dune distribution in Canada, with particular emphasis
on Alberta, Saskatchewan, and Manitoba. Although small
areas of the Great Sand Hills in Saskatchewan (David, 1993;
Wolfe et al., 1995) and the Brandon sand hills in Manitoba
(David, 1977) are presently active, eolian sands in the region
are mostly inactive. Dunes and sand sheets of the northern
Great Plains have received little attention by researchers until
recently.
Interest in eolian sand on the central and southern Great
Plains of the United States has increased dramatically in
recent years because of concerns that dunes and sand sheets in
those regions, also largely inactive, could become active
under a warmer climate in the centuries ahead, due to global
warming from greenhouse-gas emissions (Muhs and Maat,
1993). In the past, a number of investigators assumed that
dunes in the central and southern Great Plains were last active
during the last glacial maximum, ca. 22 000-16 000 BP
(Watts and Wright, 1966; Wright, 1970; Warren, 1976;
Sarnthein, 1978; Wells, 1983; Kutzbach and Wright, 1985).
Countering this concept are recent studies which show that
many dunes in this part of the Great Plains were last active
during the Holocene, and even the late Holocene (Ahlbrandt
et al., 1983; Swinehart and Diffendal, 1990; Madole, 1994,
1995;Holliday, 1995; Forman et al., 1995;Loopeet al., 1995;
Muhs and Holliday, 1995; Muhs et al., 1996, 1997a, b;
Arbogast, 1996). The conclusion that many dunes and sand
sheets in the region were last active during the Holocene,
rather than the Pleistocene, is important, because it means
that dunes can be active under climatic conditions not drastically different from those of the present.

impacts on grazing, agriculture, wetland habitats, and infrastructure. In the northern Great Plains, dune fields are of
much smaller areal extent, and direct impacts of reactivated
sand are likely to be of lesser significance. However, because
eolian sands are sensitive to changes in climate, northern
Great Plains dune fields are important as 'early warning' indicators of climate change.
The purpose of this paper is to review the work on northern Great Plains dunes conducted thus far, discuss what is
presently understood about them, and stimulate possible avenues of future research. Regional controls on sand distribution are first discussed, followed by evidence of past activity
based on geochronological, pedological, and historical evidence. It is also important to determine if present eolian activity is a function of European settlement, so criteria for this
determination are examined.

REGIONAL GEOLOGICAL SETTING,


CLIMATOLOGY, AND VEGETATION
In this paper, we define the northern Great Plains as including
southern Alberta and Saskatchewan, southwestern Manitoba,
eastern Montana, North Dakota, and the northernmost part of
South Dakota (Fig. 1). It is an area of mostly low relief, much

Recent modelling of possible dune activity in the


Canadian Prairie Provinces led Wolfe and Nickling (1997)
and Wolfe (1997) to the conclusion that significant reactivation of stabilized dunes could occur under a future greenhouse-gas-induced climate. Although Wolfe (1997)
concluded that full activation of stabilized eolian sands would
be unlikely, his results are conservative in that he used a modern analogue, the drought year of 1988, as a measure of the
degree of possible future precipitation reduction and temperature increase. Although that year had below-average precipitation and one of the warmest summers in the instrumental
record for the region, paleoclimatic data indicate that more
severe droughts are possible under a Holocene climatic
regime (Fritz et al., 1991, 1994; Vance et al., 1995).
A critical question is "What is the significance of
increased activity of presently stabilized eolian sands?' In the
central and southern Great Plains of the United States, dune
fields occupy large areas (Muhs and Maat, 1993; Muhs and
Holliday, 1995). For example, the Sand Hills of Nebraska
region alone covers approximately 50 000 krn2, and reactivation of a dune field such as this would have serious direct

Figure 1. Part of the Great Plains ofNorth America (outlined


by dashed line) and northern Great Plains (shaded), as
defined in this study. Abbreviations used: AB, Alberta; SK,
Saskatchewan; MB, Manitoba; MT, Montana; ND, North
Dakota; WY, Wyoming; SD, South Dakota; NE, Nebraska;
CO, Colorado; KS, Kansas; NM, New Mexico; OK,
Oklahoma; TX,Texas.

D.R. Muhs and S.A. Wolfe

of which was glaciated during the Late Wisconsinan (Fig. 2).


A few upland areas, such as the Cypress Hills and Wood
Mountain uplands of southwestern Saskatchewan and southeastern Alberta, were nonglaciated at that time, as was most
of the area south of the present Missouri River in Montana,
North Dakota, and South Dakota. In the glaciated part of the
northern Great Plains, drift is up to 120 m thick in parts of
Saskatchewan and Manitoba, but less than 30 m thick in other
areas, including most of North Dakota. Proglacial lakes
covered a significant portion of the area during regional
deglaciation.
The climate of the northern Great Plains is controlled by
its continental location, the rain-shadow effect of the Rocky
Mountains to the west, and the changing dominance of distinct air masses (Bryson, 1966).Due to its continental interior
setting, the northern Great Plains has long cold winters and
warm but short summers. The region is influenced by three air
masses: cold dry Arctic air; warm dry Pacific-derived air;
and, less frequently, warm moist air derived from the Gulf of
Mexico. The interaction of Pacific and Arctic air masses produces extremely strong winds that are characteristic of the
region, particularly in the western part. Winters are dorninated by cold dry Arctic air masses, such that precipitation,
even as snowfall, is minimal. Most precipitation occurs as
rain in the spring, summer, and early fall months (with a rainfall peak in June), when warmer Pacific-derived air dominates across the region. However, due to the rain-shadow
effect of the Rocky Mountains, Pacific air is relatively dry by
the time it reaches the northern Great Plains, and therefore
rainfall ranges from a maximum of only about 500 mrn.a-I in
eastern North Dakota to less than 300 mmma-l in northern
Montana, the driest part of the region. Irregular incursions of
warm moist air derived from the Gulf of Mexico reach the
southeastern part of the region, particularly southern
Manitoba and the Dakotas, and are the cause of generally
greater precipitation there. Relatively low rainfall and warm
but short summers result in an annual moisture deficit over
the northern Great Plains. Ratios of precipitation (P) to potential evapotranspiration (PE; measured by the method of

Thornthwaite and Mather (1957)) are highest (0.80-0.90) in


eastern North Dakota and southern Manitoba and lowest
(0.50-0.60) in southern Saskatchewan and northern Montana
(Fig. 2). According to the United Nations Environmental
Programme (1992) classification of arid lands, most of the
northern Great plains may be classified as subhumid to dry
subhumid.
The vegetation of the northern Great Plains reflects the net
moisture deficit. Presettlement vegetation in most of the
region was grassland. Short-grass prairie is dominant in
Montana, where the moisture deficit is the greatest (Fig. 2),
and grades to the west, north, and east into mixed-grass prairie vegetation. The mixed-grass prairie in turn grades north
into parkland, a vegetation zone characterized by
broad-leafed trees (chiefly aspen) interspersed with prairie.
Short- and mixed-grass prairie in the northern Great Plains is
bounded on the east by tall-grass prairie. Parkland on the
north is succeeded by boreal forest, and tall-grass prairie is
succeeded to the northeast by mixed deciduous and
coniferous forest. Within the northern Great Plains, local
effects of relief and aspect modify this broad biogeographic
pattern. For example, upland areas such as the Cypress Hills
are cooler and wetter than the surrounding prairie-dominated
lowlands due to local orographic effects, and therefore a
boreal forest vegetation, comprising pine, spruce, aspen, and
poplar, occurs on north-facing slopes. Deciduous forest vegetation is also found on uplands such as Moose Mountain in
southern Saskatchewan and Turtle Mountain in southern
Manitoba and northern North Dakota. Dunes themselves
often produce local variations in vegetation patterns, due to
aspect or drainage. In the Minot dune field in North Dakota
(Fig. 2), relatively cooler and moister northeast-facing slopes
of dunes support aspen and oak, whereas grass dominates the
relatively warmer and drier southwest facing slopes. In the
Pike Lake-Dundurn d u n e field of south-central
Saskatchewan (Fig. 2), moister interdune areas support
aspen, whereas grass dominates better drained dune crests
and side-slopes.

Figure 2.
Distribution of eolian sand in the northern Great
Plains region, compiled from Westin et al. (1971),
David (1977), Clayton et al. (1980), Harris (1987),
and Lord (1988),and ratio ofprecipitation to potential
evapotranspiration, based on 1961-1 990 mean values
for precipitation and temperature (computed by the
authors, wing the Thornthwaite and Mather (1957)
method). Abbreviations for dunefields and other localities referred to in text: MS. Middle sand hills; GS,
Great Sand Hills; SE, Seward; E, Elbow; PL, Pike
Lake; SL, St-Luzare; OL, Oak Lake; B, Brandon; M,
Minot; SD, Sheyenne Delta; LD, Lake Dakota; CH,
Cypress Hills; WM, Wood Mountain; MM, Moose
Mountain; TM, Turtle Mountain. Bold dashed line
shows limit of Lute Wisconsinan glaciation (Dyke and
Prest, 1987).

GSC Bulletin 534

CONTROLS ON DUNE FORMATION


Hack (1941), in studying dunes in the Navajo county of northeastern Arizona, articulated three basic requirements for
sand-dune formation: winds above the threshold velocity for
sand-particle movement, a source of sand, and a lack of stabilizing vegetation. From this basic model, geomorphologists
have developed the concepts of transport-limited dune systems (insufficient winds for eolian transport) or supply-limited dune systems (insufficient sand for significant
eolian transport). These concepts are important in evaluating
the potential for reactivation of stabilized dune systems such
as those of the northern Great Plains, since dunes may be
inactive because they are transport limited, supply limited, or
both.

Wind regimes
Wind regimes, as applied to sand dunes and sheets, are best
defined in terms of sand-moving potential, expressed graphically as sand roses (Fryberger and Dean, 1979). Fryberger
and Dean (1979) defined several parameters from sand rose
data: 1) drift potential (DP), the scalar sum of all sand-moving
winds, regardless of direction; 2) resultant drift potential
(RDP), the vector sum of all sand-moving winds (a value
always less than or equal to DP); and 3) resultant drift direction (RDD), the net direction of sand movement. Fryberger
and Dean (1979) developed a classification scheme for wind
regimes (assuming medium-sized sand particles), where DP
values of less than 200 are low-energy, 200-400 are
intermediate-energy, and greater than 400 are high-energy
regimes.

Figure 3. Comparison of mean driftpotential values for the


northern Great Plains and desert basins where large (greater
than about 30 000 km2) sand seas occur. Northern Great
Plains data computed by the authors; desert basin data from
Fryberger and Dean (1979).

Winds in the northern Great Plains are not a limiting factor for dune formation. Using records of 5-20 years and
medium sand-sized particles as an average grain size, DP values range from about 300 to about 1600 and average about
750; the overall wind regime is clearly in the high-energy category (Wolfe, 1997; Muhs et al., 1997a). In fact, average DP
values for the northern Great Plains are higher than those for
desert basins where the world's largest sand seas occur
(Fig. 3). In some parts of the northern Great Plains, seasonal
changes in wind regime (for winds above the threshold velocity for sand) occur where there are distinct changes in the seasonal dominance of large air masses. For example, in eastern
North Dakota, fall, winter, and spring are dominated by
northwesterly winds, but incursions of air masses from the
Gulf of Mexico during the summer result in the dominant
wind direction being from the south or southeast. Greater seasonal variability results in lower RDP values and limits the
potential net migration of dunes, particularly in summer
(Wolfe and Lemmen, 1999). However, compared to many
regions studied by Fryberger and Dean (1979), and even
compared to certain central Great Plains and southwestern
United States dune fields (Ahlbrandt and Fryberger, 1980;
Muhs et al., 1995, 1996), the northern Great Plains has relatively high RDP values, and net drift directions tend to be to
the northeast, east, or southeast (Fig. 4). Resultant drift
potential values are highest in southern Alberta where strong
chinook winds moderate winter temperatures.

Sources of sand
Although it is obvious that sand is a requirement for dunes or
sand sheets, the precise role of sand supply in dune formation,
degree of dune activity, and overall dune field evolution is not
well understood. Potential sand sources for dune fields
include sandstone bedrock, beach sediments, fluvial and
glaciofluvial sediments, lacustrine and glaciolacustrine sediments, and older eolian deposits. Although the sand source is
crucial to understanding how dune fields are initiated, surprisingly little effort has been made to identify sources in
most of the world's major sand seas. Dune field sources are
not always apparent, and dune fields in close proximity can
have different sources of sand. For example, the Algodones
dunes of southeastern California appear to have obvious
sources in the ephemeral streams that drain into the structural
basin where the dunes are found. However, mineralogical
data and trace element geochemistry show that shoreline sediments of a former lake, which were in turn derived from the
Colorado River, are the major source of dune sand (Muhs et
al., 1995). In northeastern Colorado, a field of parabolic
dunes is derived from sandstone bedrock while immediately
downwind, but on the opposite side of the South Platte River,
a larger field of parabolic dunes was derived from fluvial sediments (Muhs et al., 1996).
Within areas of the northern Great Plains that lie south of
the glacial limit, such as southeastern Montana and western
parts of the Dakotas, dune fields are either absent or so small
they cannot be mapped (Fig. 2). Much of the bedrock in this
area is Cretaceous or Tertiary siltstone or shale, and typically

D.R. Muhs and S.A. Wolfe

53"

112"

j; 1 -

Figure 4.

Resultant drift directions (shown by direction of


arrows) and resultant drift potential (shown by
length of arrows) compared to eolian sand distribution in the northern Great Plains. Computed by the authors using methods outlined by
Fryberger and Dean (1979).Abbreviations: CN,
Coronation; C, Calgary; SU, Suffield; L,
Lethbridge; MH, Medicine Hat; MB,
Manyberries; S, Saskatoon; 0, Outlook; SC,
Swift Current; MJ, Moose Jaw; R, Regina; CO,
Coronach; W, Winnipeg; H,Havre; G, Glasgow; B, Billings; MC, Miles City; WI, Williston;
MI, Minot; BS, Bismarck; F, Fargo.

' '

108"

~ . r , -a
, ~1

CN

\
.T
.>.,

104"

100"

96"

MANITOBA

0 7 - SASKATCHEWAN

;*.

&t

5(5s
\

'.

%G

MONTANA

NORTH *
DAKOTA.

3,
0

I
km

45O

RDP = 100

lacks sand. Furthermore, rivers are apparently not significant


sources of sand in eastern Montana or western parts of the
Dakotas. Although inany northern Great Plains rivers, such as
the South Saskatchewan, Milk, Bow, Missouri, and Yellowstone, originate in the Rocky Mountains and have
snowmelt-generated peak flows similar to central Great
Plains rivers, they are mostly deeply entrenched and therefore
do not have the channel morphologies that would be optimal
for particle entrainment by wind (cf. Muhs and Holliday,
1995).
Examination of the distribution of dune fields, proglacial
lakes, and inflow channels that created deltas in the lakes of
southern Manitoba and North Dakota supports the hypothesis
that glaciolacustrine deposits, especially sandy deltaic sediments, are the main sources of much dune sand in the northern
Great Plains (Fig. 5). In the glaciated part of the northern
Great Plains, till is areally the most important unconsolidated
sediment and does contain variable amounts of sand. However, it is unlikely that till, which contains sufficient clay to
serve as an effective binding agent inhibiting significant
eolian transport, is a significant source of sand for most of the
dune fields in the region. It is more likely that dune fields are
linked to certain facies of glaciolacustrine or glaciofluvial
deposits. David (1977) suggested that many of the dune fields
of Alberta, Saskatchewan, and Manitoba were derived from
either glaciofluvial sand or glaciolacustrine sand deposits,
particularly deltaic facies. Similarly, in North Dakota,
Bluemle (1982, 1985) and Lord (1988) hypothesized that
eolian sand in the Minot dune field was derived from density
current sediments of glacial Lake Souris, found to the northwest. Bluemle (1982) presented aerial photographs to support
the idea that the source areas in North Dakota were wind
scoured. Provenance studies of the Brandon sand hills in
Manitoba and the Minot dune field in North Dakota, using
immobile trace element geochemistry, also support this
hypothesis (Muhs et al., 1997a). Sandy facies of glaciofluvial
and glaciolacustrine deposits would have been ideal sources
for subsequent eolian transport as they tend to be moderately
to well sorted prior to entrainment by the wind.

4-

SOUTH DAKOTA

'

RDP = 200

RDP = 300

The conclusion that glaciofluvial and glaciolacustrine


deposits are the main sources for northern Great Plains dunes
means that dune fields in the region are 'closed systems' with
finite sediment supplies. In contrast, dune fields of the central
and southern Great Plains represent 'open systems', with
rivers renewing the supply of sand available for eolian transport on an annual basis (or did, prior to river regulation). In
the northern Great Plains, sand supplies are essentially relict
from deglaciation, and dune activity may be controlled more
by changes affecting the finite supply of sand available for

SASKATCHEWAN '1

1
48"

' I

CANADA
USA

Wllllston

'

MANITOBA

'L...W
Minot

G/acial Lake Agassir

Glacial
Lake
Souris

*
.---.

, ae
NORTH

n:----~*

SOUTH DAKOTA

km I p O

II

<

Glacial
Lake
Dakota

Figure 5. Map (after Muhs et al., 1997a)showing distribution


of dunefields (dark shading) in Manitoba, North Dakota, and
South Dakota in relation to proglacial lakes (light shading)
and major inflow channels (dashed arrows). Proglacial lake
and inflow channel distribution from Teller (1987).
Proglacial lake areas of eastern Saskatchewan are not
shown.

GSC Bulletin 534

transport (such as vegetation cover, disturbance, and moisture), than by replenishment of the source sands. In this
regard, dune fields of Alaska and northern Europe may be
better analogues to northern Great Plains dune Fields than
those of the central and southern Great Plains. Lea and
Waythomas (1990) argued that Holocene dune-building in
Alaska was the result of reworking of Pleistocene sand-sheet
deposits, with little or no addition of new sand. Caste1 et al.
(1989) have shown that Holocene dunes in northwestern
Europe were reworked from Late Wisconsinan sheet sands
('cover sands') as a result of human disturbances over
approximately the past 2000 years. The closed-system supply
of sediment probably explains, in part, why northern Great
Plains dune fields are smaller than those of the central and
southern Great Plains.

binding sand-sized particles and increasing cohesion, and can


also result in greater moisture retention, which further
increases cohesion.
Parabolic dunes are the dominant eolian landform in the
northern Great Plains, and a widely held concept among
eolian geomorphologists is that vegetation plays a primary
role in the genesis of this dune type (Hack, 1941; Pye and
Tsoar, 1990; Lancaster, 1995). Complete vegetation cover
will stabilize a parabolic dune, but the characteristic morphology of this dune type is generally explained by partial
vegetation cover, where the vegetation-free head migrates
downwind while partially vegetated arms either remain
behind or migrate downwind more slowly. David (1978,
1979) challenged this concept for the origin of parabolic
dunes in Canada and hvuothesized
that interstitial moisture
.
(through its effect on increasing sediment cohesion), rather
than partial vegetation cover, is the dominant factor in the
genesis of parabolic dunes. Theoretical calculations and wind
tunnel experiments by McKenna Neuman and Nickling
(1989) have quantified the relation between moisture content
and sediment cohesion. David (1978, 1979) suggested that,
were moisture content not as high as it is in the subhumid
environment of the northern &eat Plains, 'desert-type'
dunes, such as barchan dunes, would develop. It is difficult to
test this hypothesis, however, because vegetation cover on
dunes increases in part because of higher moisture content,
and both plant cover and soil moisture contents are greater in
subhumid and semiarid environments compared to arid environments. However, it is possible for 'desert-type' dunes to
develop in a semiarid climate under drought conditions and to
revert to parabolic dunes under conditions of increased
humidity. During the 1930s drought, fully active transverse
and barchan dunes developed in southwestern Kansas (Muhs
and Maat, 1993). Similarly, at the Hanford site in
south-central Washington, active sand dunes changed from
barchan, barchanoid ridge, and transverse ridged types in the
1940s to parabolic dunes from the mid-1960s, corresponding
to an observed increase in annual precipitation following
1947 (Gaylord and Stetler, 1994). Wolfe (1997) pointed out
that the White Sands area of New Mexico may be in a climatic
regime that is transitional between dunes characteristic of
semiarid and subhumid environments (parabolic dunes) and
those characteristic of arid environments (transverse and barchan dunes), because both types are presently active there.
The Monahans dune field in southwestern Texas may also be
in this category, because active barchanoid and parabolic
dunes are both present.
r

Effect of vegetation cover and moisture


Although there is general agreement among eolian
geomorphologists that vegetation cover inhibits sand transport by wind, there is little understanding of the specific processes involved (Pye and Tsoar, 1990; Lancaster, 1995).
Vegetation acts to reduce transport of erodible sediment
both by covering the surface and by trapping soil particles
(Goldsmith, 1985; Wolfe and Nickling, 1993). Vegetation
also has a direct effect on the ability of wind to transport
sediment by extracting momentum from the wind (Wolfe
and Nickling, 1993), thereby affecting the threshold velocity (Bagnold, 1941). Thus, vegetation may play a dual role
by creating both supply-limiting and transport-limiting conditions. In sparsely vegetated desert regions, the ability of
shrub vegetation to create transport-limiting conditions by
affecting the threshold velocity exceeds its ability to create
supply-limiting conditions by covering the soil (Wolfe and
Nickling, 1996). On the northern Great Plains, the much
higher density of comparatively short grass cover on sand
dunes probably has a greater impact in creating supply-limiting conditions (by covering and trapping erodible
sediment) than transport-limiting conditions (by affecting the
threshold velocity).
Vegetation cover also has an indirect effect on sediment
supply. Although plants remove moisture from dune sands
through transpiration, they also decrease direct evaporation
of moisture from dune sands through shading. Increased
moisture content in dune sands results in reduced sediment
supply, because capillary forces created between interparticle
contacts increase sediment cohesion (see discussion below).
A more subtle effect of vegetation cover is the trapping of
finer grained (silt- and clay-sized) particles. Muhs (1985)
reported silt contents of up to 20% in the surface soils of some
parabolic dunes in northeastern Colorado, and suggested that
these particles were derived from airborne dust trapped by
vegetation after stabilization of the dunes. Gile (1979), working in the southern Great Plains of Texas found that incorporation of clay and silt was significant enough to eventually
result in the development of clay-rich Bt horizons in older
dunes. Incorporation of fine-grained particles, particularly
clay, into surface soils by vegetation has the direct effect of

Dune mobility index


Because it is difficult to separate the influence of moisture
content and vegetation cover on degree of dune activity, an
alternative approach is to use climatic data that affect both
parameters. Lancaster (1988), studying dunes in the Namib
and Kalahari deserts, noted that the degree of vegetation
cover and degree of dune activity were correlated with a
regional moisture gradient. He generated an index of dune
mobility (M) that is proportional to the amount of time that

D.R. Muhs and S.A. Wolfe

wind is above the threshold velocity for sand (W), and


inversely proportional to the ratio of precipitation (P) to
potential evapotranspiration (PE). Higher values of M correspond to greater degrees of dune activity. Lancaster recognized, from field studies, four classes of dune activity:
inactive dunes (M<50), dunes with active crests only
(501Mc100), dunes that are fully active except for plinths
and interdune areas (1001M<200), and fully active dunes
(M2200).
Lancaster's (1988) index appears to have applicability to
dunes in environments other than southern Africa. Muhs and
Holliday (1995) applied the index to dunes of the central and
southern Great Plains, the Chihuahua Desert, and the
Colorado Desert of southeastern California, and found good
agreement between observed and predicted degrees of dune
activity. They reported that P P E values alone explained
degree of dune activity fairly well, suggesting that above
some minimum wind strength, degree of dune activity is
largely a function of moisture balance, either through direct
effects or through the effect of vegetation cover. Wolfe
(1997) applied the index to dune regions of the Great Plains of
Canada (his data, along with data from the central and southern Great Plains, are reproduced in Fig. 6). By both
Lancaster's (1988) original index, and Muhs and Holliday's
(1995) modified index (also shown in Fig. 6), northern Great
Plains dunes fall into the categories of either being inactive or
having active crests only, in good agreement with most field
observations (David, 1977; Running, 1996; Wolfe, 1997;
Muhs et al., 1997a). In an effort to quantify the relation
between dune morphology and climatic regime, Wolfe
(1997) suggested that the transition between parabolic dunes
and barchan dunes occurs between P P E values of 0.15 and
0.35, possibly close to 0.25. Both the Monahans dune field of
southwestern Texas and the White Sands area of New Mexico
fall within this range of the suggested transition.

80 -

~
~
active

Active
l
l except
~
for interdunes

There are some occurrences of active dunes in the northern Great Plains that are not explained by Lancaster's index.
Even though the areas are small, parts of the Great Sand Hills
in Saskatchewan and the Brandon sand hills in Manitoba
appear fully active (David, 1977,1993; Wolfeet al., 1995). In
the warmer and drier central Great Plains of the United States,
there are no dune fields that show the same amount of activity
found in the Great Sand Hills and the Brandon sand hills, and
even the southern Great Plains has mostly inactive sand north
of the Monahans dunes in southern Texas. In cooler climates
of North America, active dune fields surrounded by boreal
forest, such as the Lake Athabasca dunes of northern
Saskatchewan (David, 1977, 1989; Carson and MacLean,
1986) and the Kobuk dunes of northern Alaska (Lea and
Waythomas, 1990), are not predicted by the Lancaster dune
mobility index.
One explanation for these differences is the significantly
shorter growing season in the northern Great Plains (and even
shorter season in northern Canada and Alaska), where vegetation may take longer to recolonize dunes after major disturbances or droughts. For example, according to Wolfe (1997),
the active parts of the Great Sand Hills may be the
still-recovering remnants of dune fields that became largely
active as a result of the severe droughts of 1791-1800, based
on dendroclimatic data reported by Case and MacDonald
(1995). It is also possible that active dunes in settings such as
the Brandon sand hills and the Kobuk dunes are partially the
result of fire disturbances, such as those reported for Quebec
by Filion (1984). In the Lake Athabasca region, David suggested that the high degree of dune activity results from a
deep water table, attributed to the thickness of the sand
deposits (David, 1989) and to a comparatively short growing
season for stabilizing vegetation (David, 1979). Detailed

60 40 20 -

Inactive
0
0.0

0.2

0.4

0.6

0.8

1.0

PIPE

PIPE
Northern Great Plains

Central Great Plains

0 Southern Great Plains

Figure 6.A) Lancaster (1988)dune mobility index plot for the northern, central, and southern Great
Plains; B) modified dune mobility index of Muhs and Holliday (1995)for the same regions. W(%),
percentage of time that wind is above threshold velocidy for sand; P/PE, ratio of precipitation (P) to
potential evapotranspiration (PE). Values computed by the authors.

GSC Bulletin 534

stratigraphic studies and dating efforts of the sort conducted


by Filion (1984) are necessary to assist in determining the
causes of observed eolian activity.

Summary of controls on dune formation


Evidence from studies of dune sand provenance, eolian
mechanics, and considerations of the role of vegetation and
moisture on eolian sedimentation lead to the conclusion that
the present degree of stability of northern Great Plains dunes
is primarily a function of supply limitations. Wind strength is
extremely high, but dune activity in the region is limited
because of relatively high moisture contents and vegetation
cover, which are in turn due to the overall regional moisture
balance. The dune fields are also closed systems, with finite
supplies of glaciofluvial and glaciolacustrine sediment
deposited from the last deglaciation, which likely explains
the generally smaller size of northern Great Plains dune fields
compared to those of the central and southern Great Plains. In
addition, results obtained for much of North America suggest
that overall moisture balance (measured by P P E values),
through its supply-limiting effects stemming from both sediment cohesion and vegetation cover, is one of the more
important controls on degree of dune activity in the northern
Great Plains.

PAST DUNE ACTIVITY IN THE


NORTHERN GREAT PLAINS
Geological evidence of eolian activity
Stratigraphic and geochronological studies have been conducted at anumber of dune fields in the northern Great Plains
over the past 25 years (McCallum and Wittenberg, 1968;
David, 1971a; Wolfe et al., 1995; Running, 1995, 1996;
Burstall,
SK

Great Sand
Hills. S K

Pike Lake,
SK

Muhs et al., 1997a; David et al., 1999). Sufficient data now


exist to draw some conclusions regarding the history of northern Great Plains dune fields, although much additional work
is necessary. Studies outside the Great Plains, in northern
Saskatchewan, Quebec, and New England, demonstrate that
many dune fields record anticyclonic circulation (cf.
Kutzbach and Wright, 1985) associated with the retreat of the
Laurentide Ice Sheet (David, 1981, 1988; Filion, 1987;
Thorson and Schile, 1995). It is therefore important to
determine if northern Great Plains dunes, derived from
glaciofluvial and glaciolacustrine sediments associated with
the retreating Laurentide Ice Sheet, were last active during
deglacial time.
A compilation of radiocarbon and optically stimulated
luminescence (OSL) ages yields three important conclusions
(Fig. 7). First, the most recent periods of activity in widely
separated dune fields across the northern Great Plains were
during the late Holocene. These observations are in good
agreement with recent geochronological studies from the
central and southern Great Plains, which also document
widespread late Holocene dune activity (Ahlbrandt et al.,
1983; Swinehart and Diffendal, 1990; Madole, 1994, 1995;
Holliday, 1995; Forman et al., 1995; Loope et al., 1995;
Arbogast, 1996; Muhs et al., 1996,1997b). Documentation of
late Holocene eolian activity over such widely separated dune
fields in the northern Great Plains is important, because
paleoclimatic reconstructions for the region suggest that the
period of maximum warmth and dryness since deglaciation
was from about 7000-5000 BP (Vance et al., 1995). Over the
past few thousand years, climate and vegetation zones in the
northern Great Plains were apparently little different from
today.

A second conclusion from stratigraphic and geochronological data is that several dune fields in the northern Great
Plains have experienced alternating episodes of activity and
Brandon,
MB

Sheyenne

Minot'ND Delta, ND

Dune sand
Interdune sand

Soil, paleosol, or rhizoliths


1% age (years BP)
0 OSL age (years)

Figure 7. Stratigraphy and numerical agesfor eolian sand andpaleosols in selected northern Great
Plains dune fields. Compiledfrom data in McCallum and Wiltenberg (1968), David (1971a), Wove
et al. (1995),Running (1995,1996),and Muhs et al. ( 1 9 9 7 ~ Abbreviations
).
used: SK, Saskatchewan;
MB, Manitoba; ND, North Dakota.

D.R.Muhs and S.A. Wolfe


stability during the late Holocene. Superimposed on an essentially 'modem' climatic and vegetation regime in this region
are shorter term climatic fluctuations that can be dramatic.
Recent studies of late Holocene lake sediments in North
Dakota have documented some of these episodes. Studies
from Devils Lake, North Dakota show that there were several
episodes of relative aridity, as reflected in diatom assemblages and ostracode chemistry, in the past few hundred
years, and three major episodes of aridity in the past 1500
years (Fritz et al., 1991,1994). These interpretations are supported by studies of diatoms in sediments of Moon Lake,
North Dakota, where there is a record of as many as seven
severe droughts in the past 2300 years (Laird et al., 1996).
There are enough uncertainties associated with radiocarbon
dating of organic matter in buried soils that it is premature to
attempt correlation of the periods of eolian activity and stability with the climatic fluctuations recorded in lake cores. However, the stratigraphic data clearly indicate fluctuations in
climate over the past few thousand years, and some of these
periods of dune activity are probably due to droughts
observed in the lacustrine records.

A final conclusion that can be drawn from the stratigraphic and geochronological evidence is that periods of
dune activity and stability were not necessarily synchronous
throughout the region. In general, it appears that dune fields
on the eastern margin of the region, such as the Brandon sand
hills and the North Dakota dune fields, are more prone to
alternating periods of stability and activity, with little net sedimentation between periods of stability. In contrast, dune
fields in the more arid western part of the region, such as the
Great Sand Hills, have few or no paleosols (David, 1993;
Wolfe et al., 1995), suggesting that when reactivation takes
place, much greater volumes of sediment are involved and
evidence of previous periods of stability are removed. Given
that the western part of the region has an overall greater
potential for reactivation, based on the Lancaster mobility
index, the contrast in stratigraphic records is perhaps not
surprising.

Pedological evidence of eolian activity


Soils result from the combined factors of climate, organisms,
relief, parent material, and time. Given reasonable assumptions about constancy of the first four factors within a limited
geographic area, soils can therefore be useful indicators of
relative age. Several studies of dune fields in the Great Plains
have used soils in such a manner (Gile, 1979; Muhs, 1985;
Madole, 1995; Holliday, 1995; Muhs et al., 1996). The most
useful time-dependent soil properties are overall solum
depth; number, thickness, colour, and structural development
of B horizons; and amount of secondary clay or carbonate
buildup.

A compilation of soil data from the northern Great Plains


shows regonal differences that reflect differing degrees of
dune field activity over time (Fig. 8). In the easternmost dune
fields (Lake Dakota and Brandon), the soil landscape is a
complex mosaic of rather well developed soils with morphologically distinct Bm, Bca, or Btj horizons (Orthic Dark Grey

and Calcareous Dark Grey Chernozemic soils), and younger


soils with only Ah/C profiles (Orthic Regosols). In contrast,
dune fields farther west (Minot, Pike Lake, the Great Sand
Hills, and the Middle sand hills) show no soils with any evidence of B horizon development. Although soil B horizons
may develop faster in the moister eastern part of the region,
these observations nevertheless suggest that parts of the
Brandon and Lake Dakota dune fields have been stable for
sufficiently long periods that soil B horizons could develop,
whereas the western dune fields have been inactive, overall,
for a much shorter period of time.

Historical evidence of eolian activity


Some dune fields in the northern Great Plains, such as the
Great Sand Hills in Saskatchewan (David, 1993; Wolfe et al.,
1995) and the Brandon sand hills in Manitoba (David, 1977),
have small areas of active dunes at present. It is important to
determine if this is due to human activities since European
settlement or an indication that the dune fields are close to a
threshold of activity under natural conditions. One method of
ascertaining this is to examine records of landscape conditions by early explorers. Studies by Muhs and Holliday
(1995) using such data demonstrated that many presently stabilized dune fields in the central and southern Great Plains
were at least partially active during the nineteenth century.
One of the most detailed accounts of early explorations in
the Canadian Prairie Provinces was by Henry Youle Hind,
who travelled across parts of Manitoba and Saskatchewan in
1857 and 1858 (Hind, 1860). Hind observed active sand in
four dune fields (Fig. 9). The active parts of the Brandon sand
hills area of Manitoba were described in a colourful fashion
by Hind (1860, p. 287):
Here the sand-hills are absolutely bare, and in fact drifting
dunes.. ..In crossing the country to regain the carts, our
course lay across a broad area of drifting sand beautifully
ripple-marked,with here and there numbers of the bleached
bones of buffalo protruding from the west side of the
dunes....The progress of the dunes is very marked: old hillocks partially covered with herbage are gradually drifted by
the prevailing westerly wind to form new ones.. ..The largest expanse we saw was near the mouth of Pine Creek, it is
called by the Indians "the Devil's Hills," and a more dreary,
parched-lookingregion could scarcely be imagined...
Hind also passed by the small St-Lazare sand hills on the
Manitoba-Saskatchewan border, and it is apparent that some
of these dunes were active (Hind, 1860, p. 425):
After crossing a sandy prairie flanked on our left by
numerous bare sand hills, we reached the Assiniboine at the
mouth of the Qu'appelle [sic] early in the afternoon...
Farther west, Hind visited the Elbow dune field along the
Qu' Appelle River in Saskatchewan (Hind, 1860, p. 353):
The Sandy Hills commence on the north side about two
miles west of Sand Hill Lake as it appears in summer. They
are drifting dunes, and many of them present a clear ripplemarked surface without any vegetation, not even a blade of
grass. They have invaded the great valley, and materially
lessened its depth....Soon after breakfast we crossed the

GSC Bulletin 534

Great Sand Hills


and Middle sand hills
cm

Active
sand

Orthic
Regosol

Brandon sand hills


Active
sand

Orthic
Regosol

Orthic
Dark Gray

1: :

.....
...

Representative soil profiles found in selected


northern Great Plains dune fields. Compiled
from data in Ellis et a1 (1968). Schultz (1975,
1994), Shields and Lindsay (1988), Langman
(1989), Podolsky (1991),Acton et al. (1992), and
Muhs et al. ( 1 9 9 7 ~ )All
. soils classified according to the Canadian system of soil classflcation
(Canada Soil Survey Committee, 1978).

C : : .: .: :

U
.....

Pike Lake sand hills


and Minot dune field
cm

Orthic
Regosol

Cumulic
Regosol

C
Ahb

Lake Dakota dune field


Orthic
Regosol

Calcareous
Dark Gray

Orthic
Dark Gray

:::::

: .: : .; : .I .

I)

. 1 - *-.a;p
if.
-.t

ALBERTA

Figure 9.

108"

112O

m-

-a

104"

100"

MANITOBA

m] -

Northern Great Plains dzinefields where there is


evidence of historical dune activity. Compiled
from Hind (1860), Ellis and Shafer (1935),
Holowaychuk and Boatright (1938), Johnson
(1942), David (1977), Wove et al. (1995), and
Muhs et al. (1997~).

45"

96"

SOUTH DAKOTA

D.R. Muhs and S.A.

valley and threaded our way between sand dunes; one dune
was found to be seventy feet high, quite steep on one side,
beautifully ripple-marked by the wind, and
crescent-shaped...
Hind also journeyed north along the South Saskatchewan
River, and passed by the Pike Lake-Dundurn sand hills just
south of the present location of Saskatoon (Hind, 1860,
p. 387):
The region called the Moose Woods, which we entered last
evening, is a dilatation of the Saskatchewanflowing through
an extensivealluvial flat....This flat is bounded by sand hills,
some of which are nothing more than shifting dunes.
It is difficult to ascertain from these accounts the areal
extent of active dunes. However, these data do indicate that
some sand was active in at least four widely separated dune
fields before any major settlement of the region took place.
Thus, overgrazing or poor cultivation practices cannot
explain mid-nineteenth century dune activity. Indeed, part of
Hind's purpose in exploring the region was to ascertain
whether the lands he surveyed were even suitable for grazing
or cultivation. Therefore, a conservative but reasonable interpretation is that dunes in this region exist under a set of climatic conditions that are near a threshold between activity and
stability. This implies that only slight changes in climatic
conditions could bring about reactivation.
Several studies indicate that eolian sand was also active in
some northern Great Plains dune fields during the 1930s, the
most severe extended drought to have affected North
America during this century. Aerial photographic studies by
Muhs et al. (1997a) indicated that dunes in the Minot field
experienced partial reactivation, apparently due to climatic
forcing, in addition to human-induced reactivation of eolian
sand sheets due to a combination of cultivation and drought.
Field studies during the 1930s by Holowaychuk and
Boatright (1938) confirmed that eolian sand sheets were fully
active in many places in the Minot dune field, as their photographs show ripple marks and surfaces entirely free of vegetation. Similar studies showed cultivation-induced
reactivation in the Lake Dakota dune field during the 1930s
(Johnson, 1942). Aerial photographs also showed that dunes
in the Brandon sand hills were more active during the late
1920s and 1930s (David, 1977), but it is not known if this
greater activity was climatically forced or due to human disturbance. In the Great Sand Hills of Saskatchewan, however,
where there is no known history of cultivation, greater dune
activity during the early 1940s was probably the result of the
1930s drought, because dunes rapidly revegetated during the
wetter years of the 1940s (Wolfe et al., 1995; Vance and
Wolfe, 1996).

TIME AS A FACTOR IN DUNE


FIELD EVOLUTION
Evolution of dune fields over time has been a major concern
of eolian geomorphologists (Cooke and Warren, 1973; Pye
and Tsoar, 1990; Lancaster, 1995). In the northern Great
Plains, the combined evidence from the stratigraphy and geomorphology of the dunes, their geographic distribution, and

Wolfe

consideration of source sediments allows for some conclusions regarding the effect of time on dune field evolution.
Because almost all the dune fields in the region are within the
limit of Late Wisconsinan glaciation (Fig. 2), they are relatively young and have experienced fewer climatic fluctuations compared to dune fields in the central and southern
Great Plains, as well as those in many desert regions. Consequently, the oldest sediments are no older than Late
Wisconsinan, in contrast to some in the central Great Plains,
which are Early Wisconsinan and older (e.g. Muhs et al.,
1996). In addition, there is less evidence for complex geomorphology in the dune fields of the northern Great Plains,
with parabolic dunes and sand sheets being the main landforms, many of which are simple, rather than compound
forms. The Sand Hills of Nebraska have many complex dune
forms, such as parabolic or linear dunes superimposed upon
larger transverse forms (Swinehart and Diffendal, 1990;
Loope et a]., 1995; Muhs et al., 1997b). A further consequence of the relative youth of northern Great Plains dune
fields is mineralogical immaturity. Although the Minot dune
field lacks carbonates, thought to be due to extended periods
of eolian abrasion during the Holocene, other dune fields,
such as the Brandon sand hills, still have high concentrations
of carbonate minerals (Muhs et al., 1997a). Some central
Great Plains dune fields, on the other hand, such as the Sand
Hills of Nebraska, show a fair degree of mineralogical maturity. The Sand Hills of Nebraska even show K-feldspar depletion, suggesting a very long history that probably spans more
than one glacial-interglacial cycle (Muhs et al., 1997b).
Finally, supply limitations, as well as fewer drought cycles in
the northern Great Plains, have resulted in little growth or
migration of dune fields. Dynamic dune fields, as defined by
Pye and Tsoar (1990, p. 136), are not simply dune fields that
are active, but are those characterized by expansion and/or
migration through time. Such dune fields require not only a
lack of topographic barriers, but a continuing source of sediments. Many northern Great Plains dune fields have not
migrated far from their probable source sediments (Fig. 5),
and therefore may fall into the category of 'static' dune fields.

SAND DUNES AS PALEOENVIRONMENTAL


INDICATORS
Although much of the discussion in this review has dealt with
climatic controls on degree of dune activity, the dunes themselves, or components within dunes, can be valuable
paleoclimatic or paleovegetation indicators. Vance and
Wolfe (1996) and David (in Lemmen et al., 1998) have illustrated how dune morphology in the Great Sand Hills may be
indicative of groundwater levels and drought. Following
from their studies, David et al. (1999) dated a sequence of
dune ridges to determine eolian and hydrological events in
the Seward sand hills. This study showed that dune migration
rates, and the sequence of dune activity to stabilization, can
be reconstructed from relict dune features.
Because dunes are formed by wind, they are direct evidence of atmospheric circulation. Dune forms have particular
alignment with respect to dominant wind direction (Pye and
Tsoar, 1990; Lancaster, 1995). Parabolic dunes have arms

GSC Bulletin 534

-5.--------------------------------.
Colorado dunes
-10
Hl

---;;- -15

EB
EB

Sand Hills of
Nebraska

Minot dune field,


North Dakota

Brandon sand hills,


Manitoba

Hl

?fi

---

() -20

m~

C')

t.-0 -25
-30

500

1000

1500

2000

2500

3000

Apparent radiocarbon age (BP)


Figure 10. Carbon isotope composition of paleosol organic matter or fossil bones
from northeastern Colorado (Madole, 1995), the Sand Hills of Nebraska (Muhs et al.,
1997b), the Minot dune field (Muhs et al., 1997a), and the Brandon sand hills (David,
1971 a, b), plotted as a function of radiocarbon age. Shown for comparison are the
ranges ofvaluesfor C3 and C4 plantsfrom O'Leary (1988).
that point upwind. Odynsky (1958) measured orientations of
parabolic dunes in Alberta and reconstructed paleowinds for
the region from these data. His reconstruction showed that
dominant dune-forming winds are consistent with synoptic-scale climatology, where Pacific-derived low-pressure
cells, accompanied by strong winds, enter northern and
southern Alberta. His results are also consistent with modern
resultant drift directions shown in Figure 4. It should be
noted, however, that Odynsky's (1958) conclusions assumed
that all dunes studied are of the same age. Although soil surveys for parts of Alberta suggest that this assumption is reasonable (where all dunes have Orthic Regosols), some of the
dunes near Edmonton have Luvisols (Pawluk and Dudas,
1982), and others in northern Alberta have Brunisols.
Because Luvisols and Brunisols have Bt and Bm horizons,
respectively, these observations suggest that such dunes may
be older than those with only Orthic Regosols. A worthwhile
effort would be to stratify Odynsky's (1958) paleowind data
for each dune field, using soils as relative-age indicators.
Finally, carbon isotopes in dune paleosols can yield
important information about past vegetation types. Two types
of photosynthetic pathways, C 3 and C4 , fractionate carbon
isotopes in distinctly different ways. The C 3 plants, which
include both cool-season grasses and all trees, have 13 C values that range from about -22%o to about -34%o and average
about -26%o. In contrast C4 plants, which include most
warm-season grasses, range from about -9%o to about -20%o
and average about -12%o (O'Leary, 1988). Teeri and Stowe
(1976) found that the relative propmtion of c4 grasses in
North America is positively correlated with summer temperatures and length of growing season. The Minot dune field
and the Brandon sand hills occur in a zone of the northern
Great Plains where there was a mixture of both C3 and C4
grasses in presettlement time. A prairie with both types of
grasses will generate organic matter in soils with o13c values
that are intermediate between end member values of about

194

-12%o and -26%o. In contrast, dune fields in Nebraska and


Colorado occur in a zone of the central Great Plains where
presettlement vegetation was dominated by c4 grasses.
Radiocarbon-dated paleosols from dune fields in Manitoba,
North Dakota, Nebraska, and Colorado show ranges of values
that are consistent with the presettlement vegetation of each
subregion (Fig. 10), which suggests little difference in C 3 versus c4 grass composition at times of sand stability during the
past approximately 3000 years.

CONCLUSIONS
1. Dune fields are widespread throughout the subhumid to
semiarid northern Great Plains of southern Canada and
the northern United States.
2. Dune fields in the region are primruily supply-limited
systems. Although sand-moving winds in the northern
Great Plains are among the strongest in the world, dunes
are presently inactive because of relatively high PIPE
values, which has the dual effect of increasing moisture
content within dunes and maintaining a vegetation cover
on dunes. Dunes in the region are also closed systems, as
most were derived from glaciofluvial or glaciolacusttine
sediments deposited during deglaciation. Unlike dunes
dexived from fluvial sources in the central and southern
Great Plains, these glaciogenic deposits are finite sources,
and this probably explains, at least in part, the generally
smaller dune fields in the northern Great Plains.
3. Stratigraphic and geochronological studies indicate that
many northern Great Plains dunes are not relict features
from the last deglaciation. All dune fields studied show
evidence of late Holocene activity, and all show evidence
of having been active in the past millennium. However,
there is little evidence for regional synchroneity of dune

D.R. Muhs and S.A. Wolfe

activity, although this may be in part a function of


imprecision in dating techniques. Several dune fields
show evidence of multiple episodes of activity in the past
few thousand years. Variability in the degree of surface
soil development across the region is consistent with
these observations.
4. Historical accounts indicate some degree of dune activity
in the mid-nineteenth century, before major European
settlement of the region took place. Dunes also became
active during the 1930s drought in several regions; some
of this activity was related to cultivation combined with
drought, but other instances appear to have been driven
primarily by climate.

5. The overall picture that emerges is one of widespread


dune fields that are only marginally stable. The strong
winds and net moisture deficit combine to produce an
eolian system that is highly sensitive to small shifts in climate. The potential for reactivation of northern Great
Plains sand dunes is great, whether due to natural climatic
variations or greenhouse warming, or to anthropogenic
triggers such as those that occurred in North Dakota
during the 1930s.

ACKNOWLEDGMENTS
This study is a contribution to the Palliser Triangle Global
Change Project of the Geological Survey of Canada and the
Global Change and Climate History Program of the United
States Geological Survey. Josh Been assisted with many of
the calculations in the paper. We thank David Gaylord and
Peter David for comments on an earlier draft of the paper, and
appreciate the careful editing of Donald Lemmen.

REFERENCES
Acton, D.F., Padbury, G.A., and Shields, J.A.
1992: Soil landscapes of Canada: Saskatchewan; Agriculture Canada,
Publication 5243/B, scale 1: 1 000 000.
Ahlbrandt, T.S. and Fryberger, S.G.
1980: Eolian deposits in the Nebraska Sand Hills; United States
Geological Survey, Professional Paper 1120-A, 24 p.
Ahlbrandt, T.S., Swinehart, J.B., and Maroney, D.G.
1983: The dynamic Holocene dune fields of the Great Plains and Rocky
Mountain basins, U.S.A.; in Eolian Sediments and Processes, (ed.)
M.E. Brookfield and T.S. Ahlbrandt; Elsevier, New York,
p. 379406.
Arbogast, A.F.
1996: Stratigraphic evidence for lateHolocene eolian sand mobilization
and soil formation in south-central Kansas, U.S.A.; Journal of Arid
Environments, v. 34, p. 403414.
Bagnold, R.A.
1941: The Physics of Blown Sand and Desert Dunes; Methuen &Co. Ltd.,
London, United Kingdom, 265 p.
Bluemle, J.P.
1982: Geology of McHenry County, North Dakota; North Dakota
Geological Survey, Bulletin 74, Part I, 49 p.
1985: Geology of Bottineau County, North Dakota; North Dakota
Geological Survey, Bulletin 78, Part I, 57 p.
Bryson, R.A.
1966: Air masses, streamlines, and the boreal forest; Geographical
Bulletin, v. 8, p. 228-269.

Canada Soil Survey Committee


1978: The Canadian system of soil classification; Canada Department of
Agriculture, Research Branch, Publication 1646, I64 p.
Carson, M.A. and MacLean, P.A.
1986: Development of hybrid aeolian dunes: the Williams River dune
field, northwest Saskatchewan,Canada; Canadian Journal of Earth
Sciences, v. 23, p. 1974-1990.
Case, R.A. and MacDonald, G.M.
1995: A dendroclimatic reconstruction of annual precipitation on the
western Canadian Prairies since AD 1505from Pinusflexilis James;
Quaternary Research, v. 44, p. 267-275.
Castel, I., Koster, E., and Slotboom, R.
1989: Morphogenetic aspects and age of late Holocene eolian drift sands
in northwest Europe; Zeitschrift fiir Geomorphologie, v. 33,
p. 1-26.
Clayton, L., Moran, S.R., Bluemle, J.P., and Carlson, C.
1980: Geologic map of North Dakota; United States Geological Survey,
scale 15 0 0 000.
Cooke, R.U. and Warren, A.
1973: Geomorphology in Deserts; B.T. Batsford, London, United
Kingdom, 374 p.
David, P.P.
1971a: TheBrookdaleroad section and its significance in thechronological
studies of dune activities in the Brandon Sand Hills of Manitoba;
Geological Association of Canada, Special Paper No. 9,
p. 293-299.
1971b: Harte Road, Carberry S, and Carberry NEseries; in Geological Survey of Canada Radiocarbon Dates XI, (ed.) J.A. Lowdon,
I.M. Robertson, and W. Blake, Jr.; Radiocarbon,v. 13,p. 255-324.
1977: Sand dune occurrences of Canada: a theme and resource inventory
study of eolian landforms of Canada; Department of Indian and
Northern Affairs, Canada, National Parks Branch, Contract
NO.74-230, 183 p.
1978: Why dunes are parabolic: the wet-sand hypothesis; Geological
Association of Canada and Geological Society of America, Annual
~Meetings,
Toronto, Ontario, Abstracts with Programs, v. 3, p. 385.
1979: Sand dunes in Canada; Geos, Spring Issue, p. 12-14.
1981: Stabilized dune ridges in northern Saskatchewan; Canadian Journal
of Earth Sciences, v. 18, p. 286-310.
1988: The coeval eolian environment of the Champlain Sea Episode; in
The Late Quaternary Development of the Champlain Sea Basin,
(ed.) N.R. Gadd; Geological Association of Canada, Special Paper
NO.35, p. 291-305.
1989: Eolian processes; in Quaternary Geology of Canada and Greenland,
(ed.) R.J. Fulton; Geological Survey of Canada, Geology of
Canada, no. 1, p. 620-623 (also Geological Society of America,
The Geology of North America, v. K-I.
1993: Great Sand Hills of Saskatchewan; in Quaternary and Tertiary
Landscapes of Southwestern Saskatchewan and Adjacent Areas,
(ed.) D.J. Sauchyn; Canadian Plains Researchcentre, University of
Regina, Regina, Saskatchewan, p. 59-81.
David, P.P., Wolfe, S.A., Huntley, D.J., and Lemmen, D.S.
1999: Activity cycle of parabolic dunes based on morphology and chronology of Seward sand hills, Saskatchewan; in Holocene Climate
and EnvironmentalChange in thepalliser Triangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.
Dyke, A.S. and Prest, V.U.
1987: Late Wisconsinanand Holocene history of theLaurentide Ice Sheet;
Geographie physique et Quaternaire, v. 41, p. 237-263.
Ellis, J.H. and Shafer, W.H.
1935: Reconnaissance soil survey, south-western Manitoba; Manitoba
Soil Survey, Soils Report No. 3, 104 p.
Ellis, J.G., Acton, D.F., and Moss, H.C.
1968: The soils of the Rosetown map area 72-0 Saskatchewan;
Saskatchewan Institute of Pedology, Publication S3, 159 p.
Filion, L.
1984: A relationship between dunes, fire and climate recorded in the
Holocene deposits of QuBbec; Nature, v. 309, p. 543-546.
1987: Holocene development of parabolic dunes in the central St. Lawrence Lowland, Qutbec; Quaternary Research, v. 28, p. 196209.
Forman, S.L., Oglesby, R., Markgraf, V., and Stafford, T.
1995: Paleoclimatic significance of late Quaternary eolian deposition on
the Piedmont and High Plains, central United States; Global and
Planetary Change, v. 11, p. 35-55.

GSC Bulletin 534


Fritz, S.C., Engstrom, D.R., and Haskell, B.J.
1994: 'Little Ice Age' aridity in the North American Great Plains: a
high-resolution reconstruction of salinity fluctuations from Devils
Lake, North Dakota, U.S.A.; The Holocene, v. 4, p. 69-73.
Fritz, S.C., Juggins, S., Battarbee, R.W., and Engstrom, D.R.
1991: Reconstruction of past changes in salinity and climate using a diatom-based transfer function; Nature, v. 352, p. 706-708.
Fryberger, S.G. and Dean, G.
1979: Dune forms and wind regime; in A Study of Global Sand Seas, (ed.)
E.D. McKee; United States Geological Survey, Professional Paper
1052, p. 137-169.
Gaylord, D.R. and Stetler, L.D.
1994: Eolian-climatic thresholds and sand dunes at the Hanford Site,
south-central Washington, U.S.A.; Journal of Arid Environments,
V. 28, p. 95-1 16.
Gile, L.H.
1979: Holocene soils in eolian sediments of Bailey County, Texas; Soil
Science Society of America Journal, v. 43, p. 994-1003.
Goldsmith, V.
1985: Coastal dunes; in Coastal Sedimentary Environments, (ed.)
R.A. Davis, Jr.; Springer-Verlag,New York, p. 303-378.
Hack, J.T.
1941: Dunes of the western Navajo country; Geographical Review, v. 31,
p. 240-263.
Harris, K.L.
1987: Surface geology of the Sheyenne River map Area, North Dakota;
North Dakota Geological Survey, Atlas Series Map 15, Sheet Al,
scale, 1:250 000.
Hind, H.Y.
1860: Narrative of the Canadian Red River Exploring Expedition of 1857
and of the Assinniboineand SaskatchewanExploring Expedition of
1858; Longman, Green, Longman and Roberts, London, United
Kingdom, 425 p. (two volumes).
Holliday, V.T.
1995: Late Quaternary stratigraphy of the southern High Plains; in
Ancient Peoples and Landscapes, (ed.) E. Johnson; Museum of
Texas Tech University, Lubbock, Texas, p. 289-313.
Holowaychuk, H. and Boatright, W.C.
1938: Erosion and related land use conditions in the Minot area, North
Dakota; United States Department of Agriculture, Soil
Conservation Service, Physical Land Survey No. 3, 37 p.
Johnson, L.E.
1942: Physical land conditions in the Brown-Marshall Soil Conservation
District, South Dakota; United States Department of Agriculture,
Soil Conservation Service, Physical Land Survey No. 29, 25 p.
Kutzbach, J.E. and Wright, H.E., Jr.
1985: Siinulation of the climate of 18,000 years BP: results for the North.
AmericaniNorth AtlanticJEuropean sector and comparison with the
geologic record of North America; Quaternary Science Reviews,
v. 4, p. 147-187.
Laird, K.R., Fritz, S.C., Maasch, K.A., and Cumming, B.F.
1996: Greater drought intensity and frequency before AD 1200 in the
northern Great Plains, U.S.A.; Nature, v. 384, p. 552-554.
Lancaster, N.
1988: Development of linear dunes in the southwestern Kalahari, southern
Africa; Journal of Arid Environments, v. 14, p. 233-244.
1995: Geomorphology of Desert Dunes; Routledge, New York, 290 p.
Langman, M.N.
1989: Soils of the Rural Municipality of Victoria; Canada-Manitoba Soil
Survey, Soils Report No. D75, 149 p.
Lea, P. and Waythomas, C.F.
1990: Late-Pleistocene eolian sand sheets in Alaska; Quaternary
Research, v. 34, p. 269-281.
Lemmen, D.S., Vance, R.E., Campbell, I.A., David, P.P.,
Pennock, D.J., Sauchyn, D.J., and Wolfe, S.A.
1998: Geomorphic systems of the Palliser Triangle, southern Canadian
Prairies: description and response to changing climate; Geological
Survey of Canada Bulletin 521, p. 25-39.
Loope, D.B., Swinehart, J.B., and Mason, J.P.
1995: Dune-dammed paleovalleys of the Nebraska Sand Hills: intrinsic
versus climatic controls on the acculnulation of lake andmarsh sediments; GeologicalSociety of AmericaBuIletin, v. 107,p. 396-406.
Lord, M.L.
1988: Surface geology of the Souris River map area, North Dakota; North
Dakota Geological Survey, Atlas Series Map 4, Sheet Al, scale
1:250 000.

Madole, R.F.
1994: Stratigraphic evidence of desertification in the west-central Great
Plains within the past 1000 years; Geology, v. 22, p. 483486.
1995: Spatial and temporal patterns of late Quaternary eolian deposition,
eastern Colorado, U.S.A.; Quaternary Science Reviews, v. 14,
p. 155-177.
McCallum, K.J. and Wittenberg, J.
1968: University of Saskatchewan radiocarbon dates V; Radiocarbon,
v. 10, p. 365-378.
McKenna Neuman, C. and Nickling, W.G.
1989: A theoretical and wind tunnel investigation of the effect of capillary
water on the entrainment of sediment by wind; Canadian Journal of
Soil Science, v. 69, p. 79-96.
Muhs, D.R.
1985: Age and paleoclimatic significance of Holocene sand dunes in
northeastern Colorado; Annals of the Association of American
Geographers, v. 75, p. 566-582.
Muhs, D.R. and Holliday, V.T.
1995: Evidence of active dune sand on the Great Plains in the 19th century
from accounts of early explorers; Quaternary Research, v. 43,
p. 198-208.
Muhs, D.R. and Maat, P.B.
1993: The potential response of Great Plains eolian sands to greenhouse
warming and precipitation reduction on the Great Plains of the
U.S.A.; Journal of Arid Environments, v. 25, p. 351-361.
Muhs, D.R., Bush, C.A., Cowherd, S.D., and Mahan, S.
1995: Geomorphic and geochemical evidence for the source of sand in the
Algodones dunes, Colorado Desert, southeastern California; in
Desert Aeolian Processes, (ed.) V.P. Tchakerian; Chapman and
Hall, London, p. 37-74.
Muhs, D.R., Stafford, T.W., Jr., Been, J., Mahan, S., Burdett, J.,
Skipp, G., and Rowland, Z.M.
1997a: Holocene eolian activity in the Minot dune field, North Dakota;
Canadian Journal of Earth Sciences, v. 34, p. 1442-1469.
Muhs, D.R., Stafford, T.W., Jr., Cowherd, S.D., Mahan, S.A.,
Kihl, R., Maat, P.B., Bush, C.A., and Nehring J.
1996: Origin of the late Quaternary dune fields of northeastern Colorado;
Geomorphology, v. 17, p. 129-149.
Muhs, D.R., Stafford, T.W., Jr., Swinehart, J.B., Cowherd, S.D.,
Mahan, S.A, Bush, C.A., Madole, R.F., and Maat, P.B.
1997b: Late Holocene eolian activity in the mineralogically mature
Nebraska Sand Hills; Quaternary Research, v. 48, p. 162-176.
Odynsky, W.
1958: U-shaped dunes and effective wind directions in Alberta; Canadian
Journal of Soil Science, v. 38, p. 56-62.
O'Leary, M.H.
1988: Carbon isotopes in photosynthesis;BioScience,v. 38,p. 328-336.
Pawluk, S. and Dudas, M.J.
1982: Pedological investigation of a Grey Luvisol developed from eolian
sand; Canadian Journal of Soil Science, v. 62, p. 49-60.
Podolsky, I.G.
1991: Soils of the Rural Municipality of North Norfolk; CanadaManitoba Soil Survey, Soils Report No. D80, 128 p.
Pye, K. and Tsoar, H.
1990: Aeolian Sand and Sand Dunes; Unwin Hyman, Boston,
Massachusetts, 396 p.
Running, G.L., IV.
1995: Archaeological geology of the Rusted Quarry site (32RI775): an
early archaic site in southeastern North Dakota; Geoarchaeology,
V.10, p. 183-204.
1996: The Sheyenne Delta from the Cass phase to the present: landscape
evolution and paleoenvironment; in Quaternary Geology of the
Southern Lake Agassiz Basin, Field Trip Guidebook, Midwest
Friends of the Pleistocene, 43rd Annual Meeting, (ed.) K.L Harris,
M.R. Luther, and J.R. Reid; North Dakota Geological Survey,
Miscellaneous Series 82, p. 136-152.
Sarnthein, M.
1978: Sand deserts during glacial maximum and climatic optimum;
Nature, v. 272, p. 4 3 4 6 .
Schultz, L.D.
1975: Soil survey of Marshall County, South Dakota; United States
Governmellt Printing Office, Washinton, D.C., 116 p.
1994: Soil survey of Brown County, South Dakota; United States
Government Printing Office, Washington, D.C., 468 p.

D.R. Muhs and S.A. Wolfe

Shields, J.A. and Lindsay, J.D.


1988: Soil landscapes of Canada: Alberta; Agriculture Canada,
Publication 5237/B, scale 1:1 000 000.
Swinehart, J.B. and Diffendal, R.F., Jr.
1990: Geology of the pre-dune strata; in An Atlas of the Sand Hills, (ed.)
A. Bleed and C. Flowerday; Resource Atlas No. 5a, University of
Nebraska-Lincoln, Lincoln, Nebraska, p. 29-42.
Teeri, J.A. and Stowe, L.G.
1976: Climatic patterns and the distribution of C4 grasses in North
America; Oecologia, v. 23, p. 1-12.
Teller, J.T.
1987: Proglacial lakes and the southern margin of the Laurentide Ice
Sheet; in The Last Deglaciation of North America and Adjacent
Oceans: Responses and Mechanisms, (ed.) W. Ruddiman and
H.E. Wright, Jr.; Geological Society of America, The Geology of
North America, v. K-3, p. 39-69.
Thornthwaite. C.W. and Mather, J.R.
1957: ~nstructionsand tables for-computingpotential evapotranspiration
and the water balance; Publications in Climatoloav, Laboratow of
Climatology, centertin, New Jersey, v. 10, p. 16-3 11.
Thorp, J. and Smith, H.T.U.
1952: Pleistoceneeolian depositsof the United States, Alaska, and parts of
Canada; National Research Council Committee for the Study of
Eolian Deposits,Geological Society of America, scale 1:2 500000.
Thorson, R.M. and Schile, C.A.
1995: Deglacial eolian regimes in New England; Geological Society of
America Bulletin, v. 107, p. 751-761.
United Nations Environmental Programme
1992: World Atlas of Desertification;Edward Arnold, Sevenoaks,United
Kingdom, 69 p.
Vance, R.E. and Wolfe, S.A.
1996: Geological indicators of water resources in semi-arid environments: southwestern interior of Canada; in Assessing Rapid
Environmental Changes in Earth Systems, (ed.) A.R. Berger and
W.J. Iams; A.A. Balkema, Rotterdam, The Netherlands,
p. 251-263.
Vance, R.E., Beaudoin, A.B., and Luckman, B.H.
1995: The paleoecological record of 6 ka BP climate in the Canadian
Prairie Provinces; Geographie physique et Quaternaire, v. 49,
p. 81-98.

Warren, A.
1976: Morphology and sediments of theNebraska Sand Hills in relation to
Pleistocene winds and the developmentof eolian bedforms; Journal
of Geology, v. 84, p. 685-700.
Watts, W.A. and Wright, H.E., Jr.
1966: Late-Wisconsin pollen and seed analysis from the Nebraska Sand
Hills; Ecology, v. 47, p. 202-210.
Wells, G.L.
1983: Late-glacial circulation over central North America revealed by
aeolian features; in Variations in the Global Water Budget, (ed.)
A. Street-Penott, M. Beran, and R. Ratcliffe; D. Reidel Publishing
Co., Dordrecht, The Netherlands, p. 317-330.
Westin, F.C., Bannister, D.L., and Huron, S.D.
1971: Soil associationsof South Dakota; AgriculturalExperiment Station,
South Dakota State University, Brookings, and United States
Department of Agriculture Soil Conservation Service, Huron, AES
Information Series No. 3, scale, 1:500 000.
Wolfe, S.A.
1997: Impact of increased aridity on sand dune activity in the Canadian
Prairies; Journal of Arid Environments, v. 36, p. 421432.
Wolfe, S.A. and Lemmen, D.S.
1999: Monitoring of sand dune activity in the Great Sand Hills region,
southwestern Saskatchewan; in Holocene Climate and
Environmental Change in the Palliser Triangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.
Wolfe, S.A. and Nickling W.G.
1993: The protective role of sparse vegetation in wind erosion; Progress in
Physical Geography, v. 17, p. 50-68.
1996: Shear stress partitioning in sparsely vegetated desert canopies;
Earth Surface Processes and Landforms, v. 21, p. 607-619.
1997: Sensitivity of eolian processes to climate change in Canada;
Geological Survey of Canada, Bulletin 421,30 p.
Wolfe, S.A., Huntley, D.J., and Ollerhead, J.
1995: Recent and late Holocene sand dune activity in southwestern
Saskatchewan; in Current Research 1995-B; Geological Survey of
Canada, p. 131-140.
Wright, H.E., Jr.
1970: Vegetational history of the Central Plains; in Pleistocene and
Recent Environments of the Central Great Plains, (ed.) W. Dort, Jr.
and J.K. Jones, Jr.; University of Kansas, Department of Geology,
Special Publication, Lawrence, Kansas, p. 157-172.

Monitoring of dune activity in the Great Sand Hills


region, Saskatchewan
Stephen A. wolfel and Donald S. ~ e r n r n e n ~
Wove, S.A. and ,!,emmen, D.S., 1999: Monitoring of dune activity in the Great Sand Hills region,
southwestern Saskatchewan; in Holocene Climate and Environmental Change in the Palliser
Triangle: A Geoscientific Contextfor Evaluating the Impacts of Climate Change on the Southern
Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological Survey of Canada, Bulletin
534, p. 199-2 10.

Abstract: The Great Sand Hills and adjacent dune occurrences lie in the dry subhumid plains of southwestern Saskatchewan. Slipface advance of parabolic and blowout dunes was monitored at seven sites over
a three-year period, documenting average and maximum rates of 2.6 and 4.75 msa-l, respectively. Net
migration was eastward, but dunes may be seasonally deflected both north and south. Development and
maintenance of blowouts within stabilized dunes are the most regionally pervasive processes. Aspect plays
a critical role in blowout development, and may be positively reinforced by airflow dynamics.
The seasonal and annual variability in eolian processes is controlled by moisture availability and wind
intensity. Dunes are transport limited only during summer. Activity during the remainder of the year is
supply limited, primarily by moisture, interstitial ice, and vegetation. Greatest seasonal activity commonly
occurs in the fall, when moisture is at a minimum, seasonal vegetation is dormant, and the ground surface
has not yet frozen.

RCsumC : Les dunes Great Sand Hills et les dunes adjacentes sont situtes dans les plaines subhumides
sbches du sud-ouest de la Saskatchewan. L'avancCe des talus croulants des dunes paraboliques et des dunes
de dCflation a Ct6 surveillCe B sept sites sur une ptriode de trois ans; les vitesses moyenne et maximale
enregistrbes sont respectivement de 2,6 et 4,75m.a-'. La migration nette est vers I'est mais, selon les
saisons, les dunes peuvent Ctre dtvites vers le nord ou le sud. Le dtveloppement et la persistance de cuvettes
de dkflation au sein de dunes stabilistes est le processus le plus envahissant B 1'Cchelle rCgionale. La configuration joue un r81e critique dans le dCveloppement des cuvettes de dCflation; son influence peut &tre
renforcCe par la dynamique des flux d'air.
La variabilitk saisonnibre et annuelle des processus toliens est contr6lbe par la disponibilitt de
l'hurniditt et l'intensit6 du vent. L'activitt des dunes n'est rCgie par le transport qu'en Ctt. Le reste de
l'annCe elle est rCgie par l'approvisionnement, principalement par I'hurniditt, la glace interstitielle et la
vkgttation. L'activitt saisonnibre est gknbralement B son maximum en automne, lorsque l'hurniditk est
minimale et la vegetation saisonnibre dormante et avant le gel du sol.

Terrain Sciences Division, Geological Survey of Canada, 601 Booth Street, Ottawa, Ontario KIA OE8
Terrain Sciences Division, Geological Survey of Canada, 3303-33rd Street NW, Calgary, Alberta T2L 2A7

GSC Bulletin 534

INTRODUCTION
Wind has a significant effect on the climate and landscape of
the Palliser Triangle. Sand dunes are the most obvious landforms related to wind action, and are widespread across most
of southern Alberta, Saskatchewan, and Manitoba. However,
the relationships between climate, wind, and sand dune activity remain poorly understood. None of the available reviews
of either regional climate (e.g. Longley, 1972; Hare and
Thomas, 1979) or the eolian environment (David, 1993;
Lemrnen et al., 1998)include detailed analyses relevant to the
potential transport of dune sand by wind (Walmsley and
Morris, 1992). Furthermore, very few studies have measured
modern sand dune activity in this region (David, 1972). As a
result, our understanding of eolian sedimentary processes for
this region is based largely upon inferences from dune morphology (David, 1977).
This study presents results of a three-year dune monitoring study conducted on active blowouts and one parabolic
dune in the Great Sand Hills region of southwestern
Saskatchewan.These data help elucidate some of the dynamic
controls on dune morphology, as well as provide a baseline for
further monitoring studies into the possible effects of future
climate change and anthropogenic disturbance. Analysis of
regional wind data is presented to place the results of this
monitoring within the context of potential sand transport and
regional dune activity (see also Muhs and Wolfe, 1999).
Study area
The Great Sand Hills of southwestern Saskatchewan cover
more than 1000 km2 (Fig. I), and constitute the largest contiguous dune area in southern Canada (David, 1977, 1993).

Several smaller dune areas occur to the south and east, and
along the South Saskatchewan River to the west and north
(Fig. 1). Based on the United Nations Environmental
Programme (1992) climatic classification, the area is dry subhumid to subhumid, indicating that, although the region experiences a significant moisture deficit, true semiarid
conditions do not typically prevail. Annual potential
evapotranspiration commonly exceeds total precipitation by
more than 200 mm (Fig. 2).
All sand dunes in the region are part of the parabolic dune
association (David, 1977) and derived mainly from
glaciofluvial and glaciolacustrine deposits (David, 1964).
Although generally less than 5 m thick, eolian sands may
reach 30 m in thickness, not including the height of individual
dunes, which may be up to 15 m (David, 1964). Less than
0.5% of the Great Sand Hills is presently occupied by active
sands (Epp and Townley-Smith, 1980), with the remainder
consisting of stabilized dunes, sand sheets, and deflation
areas. Detailed discussion of the morphology and genesis of
sand dunes and associated features in this region is presented
by Wolfe and David (1997) and David (in Lemmen et al.,
1998). The terminology used in this study follows from these
papers.

Monitoring sites
Monitoring was conducted at eight sites in the Tunstall,
Seward, Bigstick, and central Great sand hills (Fig. 1). With
the exception of the 'Bowie dune', all the monitored sites
were active blowouts within stabilized parabolic dunes. The
most intensive monitoring was conducted at three sites in the
Bigstick sand hills, described in detail below.

Triangle

Figure 1. The Great Sand Hills region (dotted circle). Solid dots mark
locations of dune monitoring sites.

S.A. Wolfe and D.S. Lemmen

Bowie dune -This is an open, partially filled, parabolic dune


featuring a classic parabolic morphology consisting of a
slipface, brink, crest, backslope, and deflation area (Fig. 3).
Vegetated and partially vegetated wings occur on either side
of the active dune, whereas stabilized wing ridges extend farther upwind. The deflation depression at the base of the
backslope contains a lag of coarse sand and pebbles. Grasses
are colonizing the surface behind the deflation depression,
stabilizing the upwind portion of the dune, whereas most of
the active dune surface is devoid of vegetation. Since there
are no similarly large, active dunes in the immediate vicinity
of the Bowie dune, its activity likely relates more to local topographic effects and hydrological conditions than to modem
climate. The dune is located on a ridge of glaciolacustrine
sediments, resulting in accelerated airflow over the ridge and
locally increased depth to the groundwater table.
Baby and South dunes -These are blowout dunes located on
the heads of stabilized parabolic dunes approximately 300 m
apart. The blowouts differ in size by half an order of magnitude: Baby dune is approximately 2.5 m deep and South dune
nearly 13 m deep (Fig. 4). South dune appeared active on a
1956 airphoto, whereas Baby dune was not visible on airphotos up to 1991 and has probably formed since that time.
Despite the differences in size and age, the two blowouts are
remarkably similar in morphology. Each has a straight and

very steep-sided (30-90Q), actively eroding, north-facing


slope and a gentler, although still steep, concave-convex,
south-facing slope. Sediment cohesion is likely responsible
for the steep north-facing slopes, which are generally shaded
from direct solar insolation and retain moisture in the sand
throughout much of the year. In contrast, the south-facing
slopes are directly exposed to drying sunlight, so the surface
sands lose the cohesive effects of interstitial moisture. Near
the top of the blowouts, both slopes are near vertical (Fig. 4),
owing to the binding effects of plant roots.

METHODS
Wind analyses
In lieu of wind data from the Great Sand Hills, mean annual
potential sand transport was calculated for four locations
across southern Alberta and Saskatchewan, with Medicine
Hat and Swift Current considered most representative of the
sand hills region. Wind data suitable for determining monthly
and annual sand transport potentials were obtained from the
National Climate Archive Database of Canada. Analysis utilized mean monthly summaries for the period 1961-1990,
based on hourly data sorted into wind-speed classes by direction. From these data, sand roses were determined using the

I Ingebrigt Lake 1
30

70
60 ,
50
40
30
20 .10 g

20
w

5 10

2
$ 0

20

e
g
3 10

-s g o
E

-10
-201

30

0
-10
I

J F M A M J J A
Month

$ -10
-201

J J A
Month

Precipitation ( mm)
Location

-201

F M A M J J A
Month

70
60
50
40
30
20

10

.$

0 ;
-10
11-20
D

Climatic class.

Pot. evap.

Rein

Snow

Total (P)

(PE, mm)

P:PE

(UNEP. 1S92)

Medicine Hat

230

108

338

589

0.57

Dry subhumid

lngebrigt Lake

226

96

322

562

0.57

Dry subhumid

Swift Current

248

128

375

546

0.69

Subhumld

1 - 2 0
D

Figure 2. Monthly temperature (solid line) and precipitation (bars) summaries for three locations in the
Great Sand Hills and surrounding region. Shadedprecipitation bars denote rainfall, white bars are snowfall
in water equivalent. Data derived from 30-year normals between 1961 and 1990 (Environment Canada,
1993).

GSC Bulletin 534

Southc

..

- -

S.A. Wolfe and D.S. Lemmen


7

Figure 3.
Plan view of the Bowie dune, indicating height of the head
and backslope (at 1 m contour intervals), transect lines
monitoring the change in sulface elevation, and locations of
slipface monitoring sites.
method of Fryberger (1979). Sand roses are similar to wind
roses, but depict the potential sand transport capacity of wind
by direction, for winds above the threshold of dry sand
(>5.8 m.s-I at 10 m above ground surface in this study). Other
parameters that may be calculated include drift potential
(DP), resultant drift potential (RDP), and resultant drift direction (RDD). Drift potential indicates the monthly and annual
sand transport capacity of the wind regardless of direction.
Resultant drift potential indicates the net sand transport
capacity, based on the interaction of winds from different
directions. The ratio of RDP to DP depicts the degree of variability in the sand-transporting wind regime (where RDPJDP
=1.0 signifies unidirectional winds). Both drift potentials are
calculated in vector units (VU), which may be converted to
sediment transport rates if local material density and threshold conditions are known (Fryberger, 1979). Finally, resultant drift direction indicates the net direction of sand-moving
winds. Resultant drift direction and RDP are depicted by
arrows on sand roses, with direction reflecting the RDD and
arrow length proportional to RDP. The complete methodology for calculation and depiction of sand roses and drift
potentials is provided in Fryberger (1979).

Dune monitoring
Migration of dunes was monitored for three years (October
1993 to September 1996) at all sites except Baby dune. At
each site, the distance from the base of the dune slipface to
'fixed markers around the perimeter of the dune was measured
along fixed azimuths, documenting both migration rate and
changes in the direction of dune migration. To capture
seasonal variability, observations were made approximately
every two months for the interval October 1994to October 1995.
The net change in surface elevation of the three dunes has
been measured since May 1994, using arrays of 100 cm long
marker pins inserted 50 cm into the dune. Nineteen pins were
placed on the Bowie dune (Fig. 3), while arrays of more than
130pins were used for detailed monitoring of South and Baby
dunes. A level survey was conducted to determine initial

Figure 4.
Three-dimensional topographic representations of A) Baby
dune and B) South dune, with photographic views of the south
slopes of C) Baby dune and D) South dune. Note difference in
contour intervals in 3 - 0 representations. Note also the
stratigraphy exposed on the south face of South dune in (D)
and the steep slopes near the edge of both blowouts (areas in
shadow). Photographs by S.A. Wolfe. GSC 1999-041A,
GSC 1999-041B

elevations of each pin. Subsequent changes at each pin were


recorded by measuring the distance between the ground surface and a notch initially set 25 cm above the ground. Elevation changes were recorded at intervals of four to eight
months. When substantial erosion or deposition occurred at a
specific location, the pin was reset. In some instances, enough
erosion occurred during measuring intervals to cause a pin to
fall over. In these cases, the amount of erosion was recorded
as greater than 50 cm. In rare cases, pins were lost due to slope
failures, but this lost data is not considered to have significantly affected the long-term monitoring results.

RESULTS
Regional wind regime
Annual patterns
All four stations analyzed are characterized by a high-energy
wind regime (DP >400, Fig. 5), with potential sand transport
greater than most desert regions of the world (Muhs and
Wolfe, 1999). At Lethbridge, strong airflow off of the Rocky
Mountains results in aunimodal wind regime with low annual
directional variability (RDP/DP=0.68). Consequently, resultant sand transport is higher at Lethbridge than at localities to
the east. In contrast, Regina has the highest potential sand
transport (DP=1395), but also the greatest annual variability
(RDP/DP=0.19) and hence the lowest resultant sand transport. The bimodal distribution of winds from the northwest
and southeast at Regina may result in part from the additional
influence of air masses originating in the Gulf of Mexico
(Hare and Thomas, 1979), which is rare at the locations farther west. The resultant drift direction at both Medicine Hat
and Swift Current is northeast to east-northeast, consistent with
the orientation of sand dunes in the Great Sand Hills region.
Seasonal patterns
Although the region experiences a high-energy annual wind
regime, considerable variation occurs within the year.
Monthly potential sand transport (Fig. 6) shows a distinct
maximum in both the total sand moving capacity and variability in transport direction. Potential sand transport is at a
minimum throughout the region during July and August
(Fig. 6A). The low wind energy is primarily the result of weak
pressure gradients and convective activity associated with
thunderstorms in summer (Epp and Townley-Smith, 1980).
By contrast, potential sand transport by wind is typically at a
maximum in April and May (Fig. 6A). During autumn and
winter, potential sand transport by wind is consistently higher
than in summer months (Fig. 6A) due to more intense pressure gradients associated with frontal systems of continental
and maritime Arctic air masses (Epp and Townley-Smith,
1980). Throughout the region, wind directions are more variable during the spring and summer months (March to
September) than in autumn and winter (October to February;
Fig. 6B). However, even during the period of highest variability, net sediment transport remains toward the
north-northeast and east (025-090").

GSC Bulletin 534

Figure 5. Sand transportpotentials in southern Alberta and Saskatchewan. The arms of the "sand roses"
indicate the annual amount of potential sand transport (in vector units)from a given direction, whereas
the resultant sand transport (long arrow)for each station depicts the net trend (magnitude and direction)
of potential sediment transport. Dune areas are shaded; dotted circle outlines the same area as in Figure I.

200

. . . . Regina
- - - ,

Lethbridge

- - Swift Current
.. . . ... - Medicine Hat

Month

Month

Figure 6. Monthly potential sand transport and variability in transport direction based on hourly
depicting potential sand transport by
wind data from 1961-1990: A) monthly drift potentials (DP),
wind in vector units (VU);B) ratio of resultant drift potential to drzp potential (RDP/DP), depicting
the annual distribution in directional variability of transporting winds.

S.A. Wolfe and D.S. Lemmen

Dune migration

Sediment translocation

Annual average migration along the dune centrelines ranged


from 0.75 to 4.75 mas-l, and averaged approximately
2.6 msa-l for all sites (Fig. 7). During autumn (October to
December 1994), six of the seven dunes advanced, with the
Bowie dune and the Seward site advancing approximately
2 m during this interval, an amount that exceeded the advance
during the remainder of the 12-month period (Fig. 7). During
winter (December 1994 to March 1995), slipface advance
occurred at all locations, but at a rate considerably less than
the preceding autumn. Intercalation of sand and snow was
observed on all slipfaces, with sand occurring over snow up to
200 m downwind of active dunes. Dune migration from
spring through to mid-summer (March to July 1995) generally exceeded that during winter. The highestrates of advance
within this period occurred from March to May, but were considerably less than that which occurred the previous autumn.
During summer and early autumn (July to October 1995),
slipface retreat occurred at four of the dunes, with no significant advance measured at the other three sites (Fig. 7). Retreat
occurred in two ways: deflation by winds from v h a b l e directions, and trampling of the slipface base by cattle. Trampling
alone can result in slipface retreat of greater than l m in a season. Retreat is associated with extension of the toe of the
dune, since the sand is simply redistributed along the base of
the dune. The net result is a minor change in slipface position
in late summer.

Parabolic dune

The monitoring data also provide insights into seasonal


variability of migration direction. Figure 8A presents migration data for three points, located approximately 150 m apart
(northeast, east, and southeast lines, Fig. 3), along the slipface
of the Bowie dune. When these data are partitioned into
approximately six-month intervals, it is apparent that the
slipface was deflected both northward and southward
(Fig. 8B); however, the greatest net migration was to the east
(Fig. 8A). These short term deflections probably reflect the
inherent variability of sand transporting winds, and confirm
that interpretations of dune morphology should not be based
solely on the assumption of a unimodal wind regime.

To test the applicability of these predicted patterns, monitoring of the surface changes of the Bowie dune was undertaken for approximately two years. Results demonstrate that
the sand supply for the dune is derived almost entirely from
the backslope, with essentially no sediment contribution from
the deflation depression. Net deposition occurred between
the crest and the brink, and across the slipface (Fig. 9B). This
trend is essentially identical to that predicted by the morphological model (Fig. 9A). In addition, more than 50 cm of erosion occurred at the crest of Bowie dune over the two-year
period. The lowering of the crest is one factor contributing to
the rapid advance of the'dune slipface (Fig. 7).

Blowout dunes
The patterns of erosion and deposition at South and Baby
dunes were well developed after one year (monitoring at
three- to six-month intervals) and similar at both sites, with
erosion occurring on the eastern and southern slopes of the
blowouts and deposition occurring to both the east and west
(Fig. 10). However, maximum erosion of Baby dune
occurred along the northeastern and southwestern edges of

. . . . . . . Bigstick (Bowie dune)

Morphological models predict that migration of a parabolic


sand dune occurs as sand, derived primarily from the lower
backslope, is transported across the upper backslope and
deposited across the lee slope between the crest and slipface.
As presented schematically in Figure 9A, the dune features an
essentially flat deflation depression on the windward side
(representing a base-level substrate to which the dune has
eroded) and a backslope with a concave lower slope and
low-angled upper slope which rises to a crest. On the lee side
of the crest, the dune slopes steeply downward to the brink
(defined as the top of the slipface) and then steeply down the
slipface of the dune. As sediment is translocated, erosion is
predicted to increase downwind from the deflation depression to the top of the lower backslope and subsequently
decline across the head of the dune, reflecting the decreasing
angle of the upper backslope. Net deposition of sand is predicted
to occur beyond the crest of the dune and onto the slipface.

....." "" Seward

""""

- - - - - Bigstick (North dune)

. . GSH central (1)

Figure 7.
Cumulative slipface migration along the
centreline transect of seven monitored sand
dunes in the Great Sand Hills region. Black dots
indicate observation dates.

=>

2 4-

Oct

Jan Apr

1994

Jul Oct

Jan Apr

Jul Oct

1995

Jan Apr

1996
Year

Jul Oct

1997

GSC Bulletin 534

Figure 8.
a .

1994

1995

1996

I997

Slipface migration at three locations on the


Bowie dune: *) cumulative migration Over a
three-year period; B) incremental migration at
approximately six-month intervals.

Net
migration

Oct. Jan. Apr. Jul. Oct. Jan. Apr. Jul. Oct. Jan. Apr. Jul. Oct. Jan.

Month

I_--

Head --

I DEPOSITION (

Figure 9.

A) Cross-section of a partially filled parabolic dune depicting


major morphological features and the predicted net change
in sui$ace elevation accompanying downwind migration of
the dune (note approx. 5x vertical exaggeration).

EI

EROSION

---

B) Observed change in su$ace elevation along the centreline


of the Bowie dune (see Fig. 3),forfour surveys taken over a
two-year period (upper graph; note approx. 5x vertical
exaggeration) and the net change in sui$ace elevation over
the two-year period (lower graph).

-2

50

Mav 95

May 96
Sept. 96

100

200r

-50

150

EROSION

200
DEPOSITION

Distance along dune centreline (m)

S.A. Wolfe and D.S. Lemmen

the blowout (Fig. lOA), whereas South dune eroded primarily


along the eastern and southern walls of the blowout
(Fig. 10B). In addition, deposition on the western rimof Baby
dune occurred primarily toward the southwest, whereas at
South dune it occurred more toward the west-northwest.
Erosion of the steep, north-facing slopes of the blowouts
commonly occurs as collapse failures in late autumn or
spring, as the slopes over-steepen and blocks of sand and vegetation move downslope. Sand was most commonly transported out of the blowouts across the north and northeast
slopes by winds blowing from the west and southwest. Winds
from the east deposited sand around the western rims of the
blowouts, particularly in late summer. Since Baby dune is relatively shallow, sand can be effectively deflated from the
blowout on all sides, and deposition around the dune is primarily a function of prevailing wind direction. In contrast, the
greater depth of South dune and a sharp ridge around much of
the blowout restricts sand transport o i t of ihe blowout to the
eastern and western perimeters (Fig. 10B). The north-facing
slope is commonly either scoured or covered by slumped
sediment, indicating that it is subject to persistent erosion.
The enhanced development of the steep south slope results in
sand export to the northeast and northwest. The concave-convex profile of the south-facing slope, which is morphologically similar to the upper backslope of parabolic
dunes, suggests that sand transport dominates along this
slope. Airflow studies utilizing flagging over Baby dune
showed that, when winds approached the blowout obliquely
from the northwest, air flowed up the north slope. This suggests that a flow separation cell was set up on the north slope.
Furthermore, winds were deflected off the steeper south wall
and out of the blowout toward the northeast.
Despite significant changes along the slopes of the blowouts and deposition around their margins, the centre bases of
both blowouts showed little net change in elevation over the
monitoring period (Fig. 10). Since widening of the blowouts
occurs primarily through mass failure off the steep south
slope, there is an ongoing flow of sediment down this slope to
the base of the blowout. The sediment is then exported across
the east and north slopes. This results in only a very slow
deepening of the blowouts.

DISCUSSION
Active eolian processes
All of the climate stations analyzed are characterized by high
sand transport potential (Fig. 5). Although the directional
variability of winds increases eastward, resultant drift directions are consistently oriented toward the northeast and
southeast, a pattern that is generally reflected in the orientation of sand dunes (range 041-108"; David (1964)).
The seasonal and annual variability in dune migration is
controlled primarily by moisture availability and wind intensity. Dunes are transport limited only during summer, when
winds are weaker and directionally variable. During the
remainder of the year, dune activity is supply limited, primarily by moisture, interstitial ice, and vegetation. Greatest rates

of slipface advance occur in the fall, when moisture is at a


minimum, seasonal vegetation is dormant, and the ground
surface has not yet frozen. Although interstitial ice limits
dune activity in winter, it does not prevent it. Rather, winds
transport sand not bound by ice, which can produce an essentiall; nonerodible deflation depression and backslope until
surface sand is loosened through sublimation, melting, or
dislodgement: Although potential sand transport is greatest in
spring, directionally variable winds and abundant moisture
associated with snowmelt and rainfall limit dune migration to
less than that occurring in fall. These observations, based on
direct field measurements, appear to contrast with those
inferred by David (1993), who reported that most of the eastward migration of sand dunes occurred in late winter or early
spring. Nonetheless, the limited duration of our monitoring
must be emphasized, and it is likely that David's conclusions
are correct in years of limited snowfall and spring rain.
The mean rate of slipface advance at all monitored sites
was 2.6 m-a-l, with a range of 0.75 to 4.75 m w l . These values
are comparable to those recorded by David (1964, 1972) at
sites in the northwestern Great Sand Hills in the 1960s and
1970s, where slipface migration ranged from 0.68 to 6.47
and averaged approximately 3.6 m.a-l. In our study,
there was no obvious correlation between dune size and
migration rate. Although the largest dune monitored showed
the greatest slipface advance, this relates in part to a lowering
of the dune crest. Observations from other sites suggest that
crest lowering is fairly common and may be associated with
variations in wind strength and direction, migration of secondary bedforms, or disturbance, and can result in rapid dune
advances of up to 15 m-a-'. This is most common on dunes
with initially small slipfaces (<2 m) that often become nondiscernable following such a rapid advance.
Development and maintenance of blowouts within stabilized dunes are the most regionally pervasive active processes. Baby dune and South dune have developed on
stabilized'dune crests, sites of high localized aridity, maximum flow acceleration, and thin soil development (Hesp and
Hyde, 1996). Morphologically, dunes are comparable to
coastal dune trough blowouts: elongate, with deep deflation
basins and steep erosional lateral walls or slopes (Cooper,
1958; Hesp and Hyde, 1996). Once initiated, blowout growth
results from a positive feedback between wind flow and morphology. Blowouts in the study area feature a remarkably
similar morphology regardless of differences in age and size,
with widening and lengthening being more efficient than
deepening.
Aspect plays a key role in blowout development, maintaining differences in morphology between north and south
active slopes, primarily through differences in surface moisture. It is anticipated that air flow and topography further
influence the asymmetric morphology of these blowouts.
Results of the limited airflow observations at Baby dune are
comparable to those of Hesp and Hyde (1996), who suggested that local topographic steering may result in unequal
rates of erosion on opposite slopes of the blowouts. Therefore, the influence of aspect may be positively reinforced by
airflow dynamics in producing the observed differences
between the north and south slopes.

GSC Bulletin 534

Baby dune

South dune

Julv 94 ( 2 months)

Julv 94 (2 months)

Mav 95 (1 vearl

Distance (m)

20
0
Deposition

-20

(cm)

-40
Erosion

Distance (m)

-60

40

20

Deposition

-20

(cm)

-40

-60

-80

-100

Erosion

Figure 10. Net change in sulface elevation (erosionand depositioiz)ofA) Baby dune and B) South dune
for five intervals between May 1994 and Sept 1996. Dashed line depicts the edge of the blowout
depression at the start ofthe monitoring period.

S.A. Wolfe and D.S. Lemmen

Evolution of parabolic dunes


Relating short-term monitoring data to long-term landscape
evolution is difficult. The best data on migration rates of formerly active parabolic dunes are those of David et al. (1999)
from the Seward sand hills, and suggest mean migration rates
of approximately 2.2 m-a-l over about 60 years. These are
comparable to rates measured in this study. Examination of
regional landscapes consistently suggests that dunes have not
migrated far from their source areas (David, 1977, 1993). In
the Great Sand Hills, David (1964) estimated that dunes have
migrated a minimum of 4 km based on dune tongue morphology, whereas soil data from the western edge of same region
support an advance of 4-7 km in postglacial time (Epp and
Townley-Smith, 1980). At Tunstall sand hills (Fig. I), David
(1977) estimated Holocene migration of about 10 km.
Assuming that dunes have been active for much of the
Holocene, these data suggest long term migration rates of
0.5-2 m.a-l, also comparable to our monitoring results.
Measurement of sand movement on a fully active parabolic dune reveals patterns similar to those predicted by a
morphological model, which included a broad zone of net
sediment transport across the upper backslope, and deposition between the crest and the brink and down the slipface.
Our observations indicate that deposition between the crest
and the brink is the product of migrating bedforms and is consistent with observations of dune stratigraphy. Where
exposed in blowouts, the internal structure of stabilized parabolic dunes commonly features thick sequences of topset
beds over slipface deposits. Sedimentologically, these stabilized dunes are comparable to Type 2 parabolic dunes of
Halsey et al. (1990), in that they contain thick topset beds (>6 m)
and reflect a high sediment supply. Similarly, the Bowie dune
and other partially filled parabolic dunes in the region tend to
be relatively high, with short arms and wide heads, morphologically comparable to the Type 2 dunes of Halsey et al.
(1990). In contrast, dunes with no significant deposition
between the crest and the brink may be comparable to low
sediment-budget Type 1 dunes of Halsey et al. (1990), and the
unfilled parabolic dunes of Wolfe and David (1997).
Despite the similarities between the morphological model,
short-term monitoring, and sedimentological evidence, a morphological model of a simplemigrating equilibrium dune form
cannot be strictly applied to filled and partially filled parabolic dunes over the long term. A more appropriate model is
one that accounts for growth in dune height by aggradation of
slipface and topset deposits over time. The process by which
this occurs in parabolic dunes is still not well understood
(Lancaster, 1995).

Considerable latitudinal differences in climate exist across


the northern Great Plains, including length of the growing and
winter seasons. The comparatively short growing season of
southern Canada (less than 100 frost-free days), makes it diificult for vegetation growth to keep pace with eolian erosion and
deposition on sand dunes. Additionally, cold-climate eolian
processes (McKenna Neuman, 1993) are active in the Great
Sand Hills during winter, when low temperatures contribute to
surface drying and vegetation is dormant. This may result in
greater overall activity when compared to other Great Plain
regions with milder temperatures or shorter winters.
At a more local scale, attention needs to be given to the
distribution of active dune areas in the Great Sand Hills. Most
of these coincide with areas of high sand supply, in which the
sand is very well drained and provides a nearly unlimited supply of sediment. It is possible, therefore, that the apparently
high level of activity in the Great Sand Hills simply reflects
highly localized geological conditions. However, it must also
be noted that sand dunes in the Great Sand Hills have shown a
trend toward stabilization over the last 50 years (David, 1993;
Wolfe et al., 1995). Chronological studies by Wolfe (1996)
suggested that this trend may extend back to the early 1800s,
with the last phase of major sand dune activity being the product of a decade-scale drought in the 1790s (see also David
et al., 1999). Consequently, the high levels of dune activity
observed in a few isolated areas today may represent the last
areas to stabilize, suggesting that these dunes may not be in
equilibrium with present climate.

Implications for climate change


Monitoring the area of active sand surfaces, as well as rate
and direction of dune migration, offers insights into changes
in moisture availability and wind regimes, with implications
for human activity in the Prairies. Direct impacts of increased
sand dune activity will be local, and dune migration does not
represent a major problem on a human time scale. More significant is that an increase in the area of active sand surfaces
will result in decreased carrying capacity for cattle grazing.
Additionally, transportation corridors (roads, pipelines) and
other activities within sand hill regions could be significantly
impacted. Finally, at the regional level, sand dunes are sensitive geomot-phic indicators of climatic variability (Vance and
Wolfe, 1996), and changes in the level of dune activity may
reflect regional changes in groundwater levels and soil moisture
availability. In this respect, climatically induced increases in
sand dune activity may be a warning signal of potentially significant impacts on agricultural productivity in the Prairies.

Areas for future study


Regional dune activity
The present dune activity in the Great Sands Hills is minor in
terms of areal extent, but is greater than for most other sand
dune areas of the northern Great Plains (Muhs and Wolfe,
1999). Little consideration has been given to factors that
account for this apparently anomalous level of activity. However, several possibilities warrant discussion.

This paper presents results of dune monitoring conducted


over a three-year period as part of broader investigations of
past changes in eolian environments. As with any monitoring
study, the value of this data can only increase as the period of
observation is extended. At this stage, however, it is possible
to identify many process-oriented issues that require further

GSC Bulletin 534

research. We believe the following topics should be considered priorities for understanding dune morphology and processes in the Great Sand Hills:

1. Airflow, sediment flux, and surface soil moisture studies


should be carried out to determine the factors affecting
erosion, transport, and deposition on blowouts and
parabolic dunes.
2. Continued surface-elevation and slipface monitoring
studies should be conducted at regular (monthly) intervals
to resolve issues regarding the relative magnitude of
seasonal sand movement on dunes.

3. Stratigraphic and sedimentological investigations of


exposures within stabilized parabolic dunes should be
carried out to assess relationships between observed dune
morphology and various processes responsible for dune
deposition. In particular, the significance of sediment
supply and the relative degree of dry-sand and moist-sand
deposition should be examined.

ACKNOWLEDGMENTS
This paper is dedicated to the Bowie and the Forsyth families
who, in addition to providing access to dune areas, have cornrnunicated their strong attachment and commitment to the preservation of this sensitive environment. The co-operation of
numerous other landowners and lease holders in allowing access
to lands is gratefully acknowledged. We also thank Susan Ball,
Heather Gilmour, Kim Hodge, Lowell Strauss, and Sonya
Utting, as well as Janice Dale, Jeff Ollerhead, and David
Sauchyn, for providing valuable field assistance at various times
during the study. Review comments by Nicholas Lancaster and
Louise Filion greatly improved the presentation of this material.

REFERENCES
Cooper, W.S.
1958: Coastal sand dunes of Oregon and Washington; Memoirs of the
Geological Society of America, v. 72, 169 p.
David, P.P.
1964: Surficial geology and groundwater resources of the Prelate area
(72K). Saskatchewan; Ph.D. thesis, McGill University, Montreal,
Quebec, 329 p.
1972: Great Sand Hills, Saskatchewan; in Quaternary Geology and
Geomorphology between Winnipeg and the Rocky Mountains,
(ed.) N.W. Rutter and E.A. Christiansen; 24th International
Geological Congress, Field Excursion,Guidebook C-22, p. 3650.
1977: Sand dune occurrences of Canada: a theme and resource inventory
study of eolian landforms in Canada; Depatment of Indian and Northem Development, National Parks Branch, Contract 74-230, 183 p.
1993: Great Sand Hills of Saskatchewan:an overview; in Quaternary and
LateTertiary Landscapesof SouthwesternSaskatchewanand Adjacent Areas, (ed.) D.J. Sauchyn; Canadian Plains Research Centre,
University of Regina, Regina, Saskatchewan, p. 59-81.

David, P.P., Wolfe, S.A., Huntley, D.J., and Lemmen, D.S.


1999: Activity cycle of parabolic dunes based on morphology and chonology of Seward sand hills, Saskatchewan; in Holocene Climate
and Environmental Change in the PalliserTriangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.
Environment Canada
1993: Canadian climate normals, 1961-1990: Volume 2, Prairie
Provinces;Minister of Supply and Services, Ottawa, Ontario, 266 p.
Epp, H.T. and Townley-Smith, L. (ed).
1980: The Great Sand Hills of Saskatchewan; Policy, Planning and
Research Branch, Saskatchewan Environment, 156 p.
Fryberger, S.G.
1979: Dune forms and wind regime; in A Study of Global Sand Seas, (ed.)
E.D. McKee; United States Geological Survey, Professional Paper
1052, p. 137-170.
Halsey, L.A., Catto, N.R., and Rutter, N.W.
1990: Sedimentology and development of parabolic dunes, Grand Prairie
dune field, Alberta; Canadian Journal of Earth Sciences, v. 27,
p. 1762-1772.
Hare, K.F. and Thomas, M.K.
1979: Climate Canada; John Wiley & Sons Canada Limited, Toronto,
Ontario, 230 p. (second edition).
Hesp, P.A. and Hyde, R.
1996: Flow dynamics and geomorphology of a trough blowout;
Sedimentology, v. 43, p. 505-525.
Lancaster, N.
1995: Geomorphology of Desert Dunes; Routledge, New York, 290 p.
Lemmen, D.S., Vance, R.E., Campbell, I.A., David, P.P.,
Pennock, D.J., Sauchyn, D.J., and Wolfe, S.A.
1998: Geomorphic systems of the Palliser Triangle, southern Canadian
Prairies: description and response to changing climate; Geological
Survey of Canada, Bulletin 521,72 p.
Lonelev, R.W.
1975 - Theclimate of the prairie Provinces;~nvironment~anada,~tmos~heric
Environment Service. Climatological Studies 13.79
D.
.
.
McKenna Neuman, C.
1993: A review of aeolian transportprocesses in cold c1hal.e~;Progress in
Physical Geography, v. 172, p. 137-155.
Muhs D.R. and Wolfe, S.A.
1999: Sand dunes of the northern Great Plains of Canada and the United
States; in Holocene Climate and Environmental Change in the Palliser
Triangle: A Geoscientific Context for Evaluating the hoacts of Climate thange on the Southern Canadian ~rairiesy(ed.)D:S. Lemmen
and R.E. Vance; Geological Survey of Canada, Bulletin 534.
United Nations Environmental Programme (UNEP)
1992: World Atlas of Desertification;Edward Arnold, Sevenoaks, United
Kingdom, 69 p.
Vance, R.E. and Wolfe, S.A.
1996: Geological indicators of water resources in semi-arid environments: southwestern interior of Canada; in Assessing Rapid
Environmental Changes in Earth Systems, (ed.) A.R. Berger and
W.J. lams; A.A. Balkema, Rotterdam, The Netherlands, p. 251-263.
Walmsley, J.L. and Morris, R.J.
1992: Wind energy resource maps for Canada; Environment Canada,
Atmospheric Environment Service, Report ARD-92-003-E, 39 p.
Wolfe, S.A.
1996: Climatic, geomorphicand chnologic studies in the Great Sand Hills:
evidence for climatically induced sand dune mobilization in the last
200 years; The Canadian Association of Geographers, Annual Meeting, University of Saskatchewan,Program and Abstracts, p. 222-223.
Wolfe, S.A. and David. P.P.
1997: Parabolic dunes: examples from the Great Sand Hills, southwestern
Saskatchewan; Canadian Geographer, v. 41, p. 207-213.
Wolfe, S.A., Huntley, D.J., and Ollerhead, J.
1995: Recent and late Holocene sand dune activity in southwestern
Saskatchewan; in Current Research 1995B; Geological Survey of
Canada, p. 131-140.

Using optical dating to determine when a sediment


was last exposed to sunlight
D.J. ~ u n t l e and
~ ' Olav B. ~ i a n '
Huntley, D.J. and Lian, O.B., 1999: Using optical dating to determine when a sediment was last
exposed to sunlight; in Holocene Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate Change on the Southern Canadian
Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological Survey of Canada, Bulletin 534,
p. 211-222.

Abstract: Optical dating has now been established as a method that can be used to determine when
quartz or feldspar grains in sediment were Iast exposed to sunlight, In this review, we outline the processes
that occur in these minerals while in the environment, and the laboratory techniques used to obtain an age
estimate. The questions "Are optical ages correct?' and "What can go wrong?'are addressed. Examples are
provided of cases where optical ages are in agreement with ages obtained using other methods. Examples
are also given for instances where unexpected ages were obtained, and possible resolutions are discussed.
Emphasis is placed on the necessity of sufficient sunlight exposure prior to burial, and the degree to which
this depends on the environmental circumstances. The examples chosen are primarily from our work in
western Canada and they are intended to show what a geoscientist can and cannot expect of optical dating.

RQumC : La datation optique est dCsormais admise cornrne mCthode pouvant &tre utiliste afin de
d6terminer quand des grains de quartz ou de feldspath dans un stdiment ont 6te pour la dernibre fois exposts
2 la lumibre solaire. Dans ce texte, nous dCcrivons sommairement les processus qui se manifestent dans ces
minBraux pendant qu'ils sont dans l'environnement ainsi que les techniques de laboratoire qui sont utilisCes
pour obtenir une datation approximative. Nous abordons les questions suivantes : d e s 3ges optiques
sont-ils justes?>>et ccquelles sont les sources d'erreur?>>. Sont fournis des exemples de cas o i ~les iges
optiques concordent avec des 3ges independants et de cas oh des 3ges inattendus ont kt6 obtenus; les
possibilitCs de rtsolution des difficult& sont examin6es. On souligne la nCcessit6 que I'exposition h la
lurnibre solaire avant l'enfouissement soit suffisante et la mesure dans laquelle cela dCpend des
circonstances environnementales. Les exemples choisis proviennent essentiellement de nos travaux dans
1'Ouest du Canada et ont pour objectif de montrer ce qu'un gkoscientifiquepeut et ne peut pas attendre de la
datation optique.

Department of Physics, Simon Fraser University, Burnaby, British Columbia V5A IS6

GSC Bulletin 534

INTRODUCTION
Optical dating is a method of determining the time elapsed
since mineral sediment was last exposed to sufficient sunlight. The term 'sunlight' is used here to include any form of
daylight, whether it be direct or through clouds or water. Such
a sunlight exposure may occur between the time of erosion of
sand or silt grains and the time of their subsequent deposition
and burial in a sedimentary deposit. Some examples of processes in which such a sunlight exposure can occur are loess
deposition following deglaciation, the formation of a dune
during a dry climatic event, and the formation of a beach dune
at a high sea stand. Failure to recognize that the last sunlight
exposure is the event being dated is a common cause of difficulty in interpreting optical ages, since this event may not
coincide with the event of geological interest.
Optical dating is based on specific properties of quartz
and feldspars that depend on the existence of defects within
mineral crystals and the interaction of electrons with these
defects. When a mineral is subjected to ionizing radiation,
some electrons are ejected from their usual states and some of
these subsequently become lodged at specific defects; such
defects are referred to as traps. This is a metastable situation,
but it is possible for the lifetime of a trapped electron at ambient temperatures to exceed lo8 years. The larger the dose of
radiation, the larger the number of electrons trapped at these
defects. Traps will continue to fill until all are filled, or until
some other process leads to a state of equilibrium, or until an
exposure to sunlight or a heating event empties them.
The task of the physicist is to deduce from the trapped
electrons the magnitude of the radiation dose that put them
there and, from this in conjunction with the radiation dose
rate, to calculate an age. The age of the sample is calculated
using the formula:
Age =

Radiation dose
Radiation dose rate

Optical dating is one of three techniques based on this


principle; the other two are thermoluminescence dating and
electron-spin-resonance dating. The methods do not overlap
as much as one might imagine, because each generally makes
use of different defects.
Radiation dose is strictly defined as the amount of radiation energy that is absorbed per unit mass of the mineral; the
SI unit is the gray (Gy), with 1 Gy being equal to 1 J.kg-l.
However, the actual radiation dose is not what is determined
in any of these dating techniques. This is because the radiation dose arises not only from cosmic rays, but also from
alpha, beta, and gamma radiation caused by the decay of
potassium, thorium, uranium, and their daughter products
within the sediment and its surroundings. As well, the number
of trapped electrons depends not only on the dose but on the
type of radiation. For instance, alpha-particles are only about
one-tenth as efficient, on a per-unit-of-absorbed-energy
basis, as the other forms of radiation. To deal with this, the
term 'equivalent dose' was introduced; it is defined as the

dose of beta or gamma radiation that would produce the same


effect as the actual radiation dose (i.e. thatreceived in the natural environment). It is commonly abbreviated D, or Deq.
One can distinguish conceptually among the following
quantities: a sequence of dose intercepts on graphs made from
data obtained in the laboratory; an estimate of the equivalent
dose, deduced from those intercepts, which is used to calculate the age; and the actual equivalent dose. There is no
accepted terminology that distinguishes among the three, and
the term equivalent dose is often used to refer to any of them.
Quartz and feldspars are used for optical dating because of
their ubiquity and the ease with which they can be separated
from sediment, and because they yield ample luminescence
for measurement. Other minerals can be used in principle, but
there has been little research done on them.
Any sediment that contains quartz or feldspar and has
been exposed to sufficient sunlight prior to burial is suitable
for optical dating. What is 'sufficient' sunlight exposure is
discussed in some detail later in this paper. Depositional environments most likely to satisfy the requirement are eolian,
such as sand dunes and loess, and some low-energy water
environments, such as peat bogs. In order of decreasing likelihood, we would add lacustrine, fluvial, and glacial environments. Such a list can be misleading, because there is loess
that is not suitable and there are some glacial deposits that are;
the key is the amount of sunlight exposure to which the grains
were subjected. The established age range for optical dating
is from about 1000 to 100 000 years. What can be achieved
depends very much on the minerals present in the sample and
their behaviour. Ages of 100 years or less can readily be
obtained, and may be credible if proper techniques are used.
For some samples, both quartz and feldspar have sufficient
sensitivity to produce ages of the order of one year, but there
are presently no reliable techniques for achieving such an
age. 'Ages' on the order of 1000 000 years have also been calculated, but again reliable techniques for obtaining accurate
ages of this order have not been developed.
Problems that can occur with optical dating are discussed
in detail later. Here we note that, as with other dating rnethods, one should put little faith in an isolated optical age.
Instead, dates should be obtained for a suite of samples that
are part of a detailed research program, and those dates should
be considered in the context of the overall picture. This is
because one's intuition about the adequacy of sunlight exposure at the time of deposition is often wrong, and the laboratory techniques in use are not always appropriate.
Earlier reviews of optical dating have been published by
Aitken (1992, 1994) and Berger (1995). Sediments can also
be dated using therrnoluminescence (Berger, 1988,1995). In
most cases, optical dating is simpler and superior. The one
notable exception is that quartz that is too old for optical dating can sometimes be dated using thermoluminescence. Optical dating is simpler because heating in a controlled
atmosphere is replaced by an exposure to light in air. It is
superior because less sunlight exposure is required before
burial, the traps that are most easily emptied by sunlight are

D.J. Huntley and O.B. Lian

intensity of this light emission (luminescence) is thus a measure of the number of electrons trapped since the last sunlight
exposure. The longer the mineral grains remain shielded from
sunlight, the more light they will emit when measured.

sampled preferentially (leading to the ability to obtain much


younger ages), and single-aliquot methods that make use of
much less sample are feasible. In addition, advantage can be
taken of the ability to select the energies of the excitation photons in order to target particular minerals or defects.

The measurement procedure simply involves shining


light on mineral grains separated from the sample (Fig. 1A)
and measuring the light that is emitted in response (Fig. 1B).
In each case, the light should be thought of as a stream of photons, their energies being El for the incident beam, and E2 for
the emitted photons. The photons of the incident beam need
not necessarily all have identical energies, but for practical
reasons only a small range is used. The emitted photons may
have a relatively large range of energies, spread over one or
more emission bands that depend on the nature of the minerals present. Two of the most common combinations of El and
E2 that are used in practice are shown in Table 1.The emitted
photons are counted using a photomultiplier tube, the output
of which is an electrical pulse for each photon detected, and
some electronics that count these pulses. Figure 1B shows

Some of the terminology used varies from author to


author, so the meanings of some of the more common terms
found in the literature are included in Appendix A.

MEASUREMENTS
Sunlight ejects electrons from traps, and environmental radiation after burial slowly puts electrons back into the traps.
Exposure to light in the laboratory ejects the electrons which,
having an excess of energy, may lose it by emitting light. The
last process will occur at a particular type of defect that is
referred to as a luminescence or recombination centre. The

Table 1. Commonly used incident and measured photon energies;


wavelengths and colours are given in parentheses.

A
lncident photon beam (E,)

2.0

Sample CCL3
a-OGyy

b-12Gyy
C-25Gyy
d - 37Gy y
e-62GYy
f - 3 h bleach

04

20

Incident photon
beam on

40

60

Time (s)

80

1004

120

Incident photon
beam off

Figure 1. Principles of measurement in optical dating. A) A beam ofphotons of energy El is shone on


grains separatedfrom the sample; these grains respond by emittingphotons of energy E2. The longer
the sample has been buried, the greater the number of emitted photons. B) The decrease in
luminescence intensity with time, as the traps are emptied by the laboratory excitation. The different
curves arefor aliquots given the radiation doses shown, and arefor4-11 pm grainsfrom a 2.7-2.4 ka
loess. Curve f is from an aliquot that had received a red-infrared bleach, obtained using a
quartz-halogen lamp and long-passfilter with 50% transmission at 715 nm. Each curve is an average
of several determinations. More data for this sample are shown in Figure 2 (cfi CCL3 in Table 2).

GSC Bulletin 534

typical curves of count rate (luminescence intensity) versus


time, the decrease in luminescence occurring as the traps are
emptied.
Because a large number of the incident photons are simply
scattered by the grains of the sample, while other processes
lead to emission of photons at different energies (ordinary
photoluminescence and Raman scattering in the grains), one
measures only photons of significantly higher energy than the
incident photons (Table 1).Optical filters are used in the measuring system to ensure that a sample that has just been well
exposed to sunlight will not scatter or emit any photons in the
energy range being measured. In the case of quartz, the small
amount of information available indicates that there is little
choice of energy and filter combinations: there is only one
emission band and incident photons of different energies are
all emptying the same traps (Huntley et al., 1996). Feldspars,
on the other hand, offer more options because there are several emission bands and traps. However, the potential for
using these options has not yet been fully explored. There is
also the option of employing inclusions, probably feldspar,
within quartz grains (Huntley et al., 1993). Indeed, any given
sample may present several options for optical dating.
The actual apparatus used for the measurements can be
quite simple, such as that used in the initial work (Huntley
et al., 1985). The measurements lend themselves readily to
automation and there are at least three commercial instruments available; one of these is described by Bfltter-Jensen
and Duller (1992) and Markey et al. (1997). Such instruments
can be very useful for routine work, but they sometimes lack
the flexibility needed for more fundamental research and
their 'black-box' nature allows novice users a broad scope for
making errors. For a sample that yields little luminescence, it
is worthwhile measuring as many of the emitted photons as
possible; at present, the most efficient apparatus in this
respect is the design of Baril(1997). Instruments for field use
are possible (Poolton et al., 1994) and may be valuable for
comparative work, but they should not be expected to provide
ages.

after the laboratoly irradiation; these electrons are not present


after an environmental radiation dose, which occurred over a
much longer period of time. There is no consensus on what
this treatment should be, but five minutes at 220C, 16 hours
at 120C, or seven days at 140C are typical; the treatment
chosen is based on convenience as much as science. It is clear
that the temperature and duration used may make a difference
and, in at least one instance, there is a strong difference of
opinion on the matter (Roberts et al., 1994; Stokes and
Gaylord, 1994).

-20

20

40

60

Dose (Gy)

Sample CCL3

DETERMINATION OF THE
EQUIVALENT DOSE
The central problem here is to determine the laboratory radiation dose that produces the same population of trapped electrons as did the environmental radiation dose. The methods
used can be divided into two groups: additive dose and
regeneration.
The additive-dose method is illustrated in Figure 2. Here,
aliquots consisting of grains separated from the sample are
given different doses of laboratory radiation, heated, stored,
and their luminescence measured; a line is then fitted to the
plot of luminescence intensity versus laboratory radiation
dose. The line is commonly extrapolated to zero intensity,
and the dose intercept taken to be the equivalent dose. However, for reasons described below, this procedure is not correct, although the error introduced may not be significant for
relatively old samples. The heating is required to de-trap
some thermally unstable trapped electrons that are present

0 0 0-

20

40

60

80

100

Time (s)

Figure 2. A) Illustration of the additive-dose method with the


thermal transfer correction. The intensity of the
luminescence resulting from optical excitation is plotted as a
function of laboratory radiation dose for two sets of aliquots.
The photon counts shown are for thefirst 10 s of excitation.
The solid circles are 'additive-dose' data from aliquots that
have been irradiated and heated. The open diamonds are
from aliquots that have been irradiated, given a light
exposure, and heated, and are used to correct for thermal
transfer. The equivalent dose is taken to be the radiation dose
where the two lines intersect (marked by the arrow), with a
small correctionfor decay during the normalization. B) Dose
intercept determined as in (A), as a function of illumination
time. Also shown is the intensity versus time curve for an
aliquot that has only been preheated in order to show the time
scale.

D.J. Huntley and O.B. Lian

Unfortunately, this heating has an unwanted effect. This


is most readily observed by taking grains that have just been
exposed to sunlight, heating them, and then measuring them.
Without the heating no luminescence is measured, whereas
with it a substantial amount may be measured. The reason for
this is that, at high temperatures (e.g. 160C), electrons are
emptied from traps that are not emptied at the environmental
temperature of approximately 20C, even over a long time
period. Some of these electrons are subsequently retrapped in
traps being sampled during the measurement process. This is
referred to as 'thermal transfer' and it can be corrected for as
shown in Figure 2. The number of electrons in the source
traps for the thermal transfer depends on the radiation dose, so
it is necessary to make the thermal transfer measurement as a
function of dose. Therefore, several of the aliquots are
exposed to laboratory light to empty the traps of interest and
then heated along with the ones that received no laboratory
light. The luminescence measured from the former defines
the lower data set shown in the figure, and the dose intercept
of the two lines is taken to be the equivalent dose. If this correction is not made, then a substantially incorrect age can be
obtained, particularly for young samples. In the example
(Fig. 2), we show data obtained for loess directly overlying
Bridge River tephra, which is known to have been deposited
2700-2400 years ago (Clague et al., 1995; Leonard, 1995).
The optical age obtained using the thermal transfer correction, 2500 f300 years, is in excellent agreement. Had the dose
axis intercept been used, an erroneous age of 4000 years
would have been obtained instead. This method of making
the thermal transfer correction also allows for the dark count
rate from the photomultiplier tube and scattered excitation
photons that are counted. In the example shown in Figure 2,
all three effects were significant.

laboratory light exposure. There is some concern that the laboratory light exposure may have some unwanted effects, such
as a change in the dose response of the sample; in practice,
some test should be made to determine whether or not the two
data sets do actually define a single curve. The duration and
spectrum of the laboratory light used may well be important
here. Sunlight is often used because that was what was present at the time of deposition and the original sample condition
should be reproduced as closely as possible. However, this
argument does not withstand scrutiny, since the original
sample condition may not be obtained and may not even be
attainable. For K-fel.dspars, we have been using an infrared
exposure, which typically reduces the measured 1.4 eVexcited luminescence intensity to about 3% of the original
with no apparent detrimental effect.
With either the additive-dose or the regeneration method,
the equivalent dose should be determined as a function of illumination time. It should be constant, as shown in the example
in Figure 2B. There is cause for concern if it is not constant, a
situation which is not yet understood but which may arise
from inadequate sunlight exposure before burial.
With both additive dose and regeneration methods, scatter
results from the fact that all the aliquots prepared from a sample are not identical. This is due primarily to the grains being a
highly inhomogeneous collection, with most of the measured
luminescence arising from a small fraction of them. A
method that is in common use for alleviating this problem is
to initially make a brief measurement of each aliquot and to

TAGLU core (22.0 m)

We have found a number of quartz samples for which the


luminescence from thermal transfer so overwhelms what we
are trying to measure that it has defeated our attempts at optical dating. For some feldspar samples that are older than ca.
1000 years, we have evidence that the above method of correcting for thermal transfer is not quite right, but the reason
for this is not yet understood. We have also found that the
actual laboratory light exposure used to make the thermal
transfer correction can be important (Huntley and Clague,
1996).
The additive dose method is quite suitable when the luminescence versus dose data are linear, or nearly so, and an
accurate extrapolation can be made. When this is not the case,
a regeneration method should be seriously considered
(Fig. 3). Here, one group of aliquots is given a variety of radiation doses, while a second group is exposed to laboratory
light to empty the traps and then given similar radiation doses.
The second group, after the light exposure, resembles hypothetical aliquots prepared immediately after burial. Both sets
of aliquots are then heated and measured. The data from the
second group of aliquots are referred to by the term 'regeneration', and may be thought of as defining the missing portion
of the additive dose curve that would otherwise have to be
extrapolated. The equivalent dose is taken to be the shift
along the dose axis that brings the two data sets into alignment, with a correction for the incomplete effect of the

500

1000

1500

Dose (Gy)
Figure 3. Illustration of the regeneration method. The two
data sets are an additive-dose set (solid circles) and a
regeneration set (open squares). The latter are for aliquots
that have been given a light exposure, a laboratory radiation
dose, and then heated. The additive dose data have been
shifted horizontally to give the bestfit. The equivalent dose is
taken as the shift, with corrections for incomplete trap
emptying by the laboratory light exposure and decay
resulting from the normalization measurement. The data are
for a sediment from the TAGLU core, Mackenzie River delta
(Huntley, 1997), and are for the first 10 s of excitation.

GSC Bulletin 534

subsequently use this luminescence to normalize the data. If


this measurement empties a significant portion of the traps,
say 5%, then it must be corrected for when the equivalent
dose is calculated.
The uncertainty in the equivalent dose is obtained using
proper statistical techniques in fitting thedata; it is usually the
dominant contribution to the uncertainty in the final age.

DOSIMETRY
The radiation dose results from the radioactive decay of K, U,
Th, Rb, and their decay products, both from within the sediment grains and from the surroundings, up to about 50 cm
away. There is also a contribution from cosmic rays. All of
these have to be evaluated. The various contributions to the
dose rate (strictly the equivalent dose rate) are shown in
Figure 4 for a typical K-feldspar grain in a sample collected
from a sand dune.
Dose rates are typically in the range 2-3 Gy.ka-l, but can
be up to five times smaller or larger. Because the dose rate is
the sum of several contributions, what might normally be
thought of as a large uncertainty in one of them, say f lo%,
will lead to a much smaller uncertainty in the final age. There
is little merit in striving for high accuracy in any one term,
with one exception: the evaluation of the beta dose rate from
K and Rb within K-feldspar grains. The uncertainty in the K
content of the K-feldspar grains that contribute to the
luminescence being measured may well be f20%, and this
can lead to an uncertainty in the age of between +2 and
f 10%.
Water in the sample matrix reduces the dose rate. If the
water content is known, its effect can be readily calculated.
What can be difficult is estimating the history of the water
content throughout the existence of the deposit, something
Cosmic rays

( ~ o t adose
l
rate = 2.69 ~yka-'l

Figure 4. Illustration of the various contributions to the dose


rate. The example isfor a typical 250,um diameter K-feldspar
grain at 1 m depth in a sand dune consisting of homogeneous
material containing 1% K, 1,ug.g-' U, 3 ,ug.gl Th, and 5%
water. The K-feldspar grain has 13% K, 250 ,ug.g-l Rb,
0.2 ,ug.g-] U, and 0.4 ,ugug;g-l
Th.

that must be done by the geoscientist. Sometimes, the best


that can be done is to estimate the two extreme possibilities
and use them for the 95% confidence limits. In cases where a
sediment has been water saturated for a portion of the time
and wet to dry for the rest, the uncertainty in the final age may
be dominated by the uncertainty in the average water content,
and may be as high as d o % .
Evaluation of the dose rate can be made using laboratory
analyses for K, U, Th, and Rb contents, thick-source alpha
counting, beta counting, or high-resolution alpha or gamma
spectrometry; in situ gamma spectrometry and thermoluminescence dosimetry may also be used to determine the gamma
dose rate. More than one set of measurements should be used if
high accuracy is desired or if the sediment is inhomogeneous
within 50 cm of the sample; in this way, difficulties associated with inhomogeneity or radioactive disequilibrium can be
identified and allowed for. Finally, we should point out the
necessity for vigilance: four of the five laboratories we have
employed to do analyses have on occasion provided us with
significantly incorrect values.

SAMPLE COLLECTION
Three things must be kept in mind during sample collection:
1) the grains that are to be measured must not be exposed to
light; 2) samples of differing material will be required and the
geometry must be recorded if the sediment is not homogeneous within 50 cm of the sample in all directions (in situ
dosimetry is an alternative and often preferable method); and
3) estimates of present and past water contents will be needed.
Sample collection is best done by someone from the optical dating laboratory. If this is not feasible, close collaboration is essential. If accuracy is a major concern, then in situ
gamma spectrometry should be performed and personnel
from the dating laboratory may be able to do this.
About 1kg of material is usually more than adequate, and
should be collected if it is available. There are occasions
when 1 kg is not adequate and other cases where it may be
possible to obtain a date from a sample as small as 1 g; the
details are numerous and will not be discussed here. If the
sediment is loose, then a fresh face should be exposed and the
sample placedin a 1 L lined can and sealed; this is best done at
night, although it can be done quickly under a black cloth in
the daytime if necessary. Material that can be carved into a
block can be collected in daylight, wrapped in foil, and
sealed. The outer layer is subsequently removed from the
block in the laboratory and discarded. Samples can also be
collected by pushing a tube into the sediment. The
geoscientist should provide the laboratory with information
on the sample material, such as grain-size distribution,
organic content, and carbonate content, and find out from the
laboratory exactly what is possible.
In general, sediment that has not been collected with optical dating in mind will not be suitable. There are exceptions,
but they are rare and reduced accuracy should be expected.

D.J. Huntley and O.B. Lian

Table 2. Selected successful comparisons of optical ages with independent ages.


Independent
mlneral

Optical

Materlal

1.4 eV fg
LCPS
MPSl

Peat2
Peat2
Organic-rich
floodplain silt
Loess

18 000-1 9 000
25000-32000

16400*1500
31 200*2600

Loess
Tsunami-laid sand
Dune sand
Dune sand

'

I SESA-71

Dune sand

I Dune sand

Calendar ages before AD 1995.


Inorganic silt fraction was dated.
Mineral abbreviations: fg, 4-1 1 pm, polymineralic silt grains; Kf, sand-sized K-feldspar grains;
qi, feldspar(?) inclusions within sand-sized quartz grains; qz, sand-sized quartz grains.
References: 1, Lian et al. (1995); 2, Lian (1997); 3, W.J. Vreeken, D.J. Huntley, and S.A. Wolfe
(unpub. data, 1997); 4, Huntley and Clague (1996); 5, Huntley et al. (1993); 6, Huntley et al. (1996).

ARE OPTICAL AGES CORRECT?


This is not an easy question to answer. The basic physics is
understood in outline and a number of tests have given ages in
agreement with those provided by other methods. On the
other hand, the physics is not understood in detail, and even
the identities of most of the relevant defects are not known.
Moreover, there are effects observed that have yet to be
explained and incorporated into the methodology. This is further complicated by the fact that the mineral grains being
measured are usually a highly inhomogeneous collection,
even within a single mineral type from the same source area.
The first recourse is to perform independent tests of the
validity of optical ages. It is highly desirable in any suite of
samples to have some checks. These should include a modern
(i.e. zero-age) analogue, although admittedly one can never
be sure that one has sampled a true modern analogue; if the
age obtained for it is not zero, then something may well be
amiss. Other checks may be provided by radiocarbon or other
dating methods, or by volcanic ash or other markers of known
age. Paleoclimatic information can be used when it can be
related to an oxygen-isotope interval of known age. Table 2
shows some selected successful comparisons of optical ages,
determined at the SFU laboratory, with ages obtained from
other methods. Additional comvarisons can be found in
Smith et al. (1990), Stokes and daylord (1993), and Berger
(1995).
Samples HCPl and LCPS are compressed peat deposited
during the Port Moody interstade (6180 stage 2) and the
Olympia nonglacial interval (6180 stage 3), respectively,
while sample MPSl is an organic-rich floodplain silt
deposited during 6180 substage 5e; all three are from southwestern British Columbia. Samples CCL3 and CCL4 were
taken immediately above and below a layer of Bridge River
tephra in a postglacial loess deposit in south-central British
Columbia. Sample SAW94-62 is a postglacial loess, containing Glacier Peak tephra, from the Cypress Hills, Alberta.

Sample CBTS2 is a tsunami-laid sand from Cultus Bay, near


Seattle, Washington. The SESA samples are from beach
dunes, in southeastern South Australia, which formed during
high sea stands.
Table 2 shows selected results, and should not be used as a
general guide to the validity of any particular optical age. It
does show, however, that valid optical ages can be obtained
with appropriate techniques, and that an accuracy of 10% or
better is attainable. It is quite possible to obtain a precision of
5% or better, but it is unlikely that the accuracy will be this
good because of unknown changes in the radiation dose rate,
mineral behaviour that is not allowed for in the equivalent
dose determination, and inadequate sunlight exposure prior
to burial.
A second recourse is to obtain optical dates using different
minerals or different excitation energies to sample different
defects. The little work that has been done in this area usually,
but not always, shows agreement between the two ages;
examples can be found in Questiaux (1991), Huntley et al.
(1993, 1996), Wolfe et al. (1995), and Rees-Jones and Tite
(1997). One example is given in Table 2 (sample SESA-7 1).

WHAT CAN GO WRONG?


Things that can go wrong can be divided into two areas, those
that are in the domain of the field scientist, and those that are
in the domain of the dating laboratory.
Probably the most important thing that can go wrong is
that the sample was not adequately exposed to sunlight prior
to the last burial. If the sample was not exposed at all, then an
age corresponding to an earlier exposure event will be
obtained; if the sample was partially exposed, then an intermediate age will be obtained. It is essential that the
geoscientist has a clear picture of mode(s) of transport and
deposition. Unfortunately, knowledge on this point is often

GSC Bulletin 534

limited and intuition is often wrong. If there is error in this


regard, it is usually that there was much less sunlight exposure than envisaged. For this reason, it is essential that the
geoscientist, if at all possible, collect a modern analogue (i.e.
a sample of similar mineral composition, derived from the
same source area, and very recently deposited under analogous conditions). The geoscientist is also responsible for providing information about the past water content, possible
leaching, or other changes in composition of the sediment
with time, and the likelihood for disequilibrium in the uranium decay series, since all of this is required for calculation
of the dose rate. For example, disequilibrium will occur if any
of the isotopes in the 2 3 8 decay
~
chain move out of, or are carried into, the sediment; some examples have been given by
Prescott and Hutton (1995).
There are many possibilities for error in the dating laboratory. The proper use of laboratory controls and alertness to
potential problems are essential, since optical dating cannot
yet be considered a routine procedure. Different laboratories
use different techniques for the evaluation of the equivalent
dose. In some cases, the differences are unimportant; in others, the technique used may be invalid. Even the most accomplished laboratories may generate incorrect values because of
mineral behaviour that is not recognized as being anomalous.
It is difficult for a nonexpert to evaluate this, although there
are some tests and considerations that can be applied:
1. For relatively young samples, a thermal transfer
correction must be made; if this is not done, any age
calculated is likely to be too old. It is difficult to be
quantitative about the effect of this correction: one
example has already been given while another is that of a
300-year old sample for which an age of 4500 years would
have been calculated if the thermal transfer correction had
not been included (Huntley and Clague, 1996).

5. Plots of dose intercept versus illumination time should be


shown and there should be an explanation of how
equivalent doses were obtained from them.

6. An understanding of the scatter in the laboratory data is


desirable; in some cases, this can reveal inadequate sunlight exposure before burial.

HOW MUCH SUNLIGHT EXPOSURE


IS SUFFICIENT?
The answer to this question depends on the circumstances, so
the question is best addressed by discussing examples.
Figure 5 shows typical effects of sunlight exposure on quartz
and feldspar grains cleaned and separated from deposits that
are ca. 500 000 and 70 000 years old, respectively. If the
70 000-year old deposit was eroded, exposed to sunlight,
redeposited, and buried, with all the feldspar grains receiving
500 seconds of the same sunlight that was used to obtain the
data of Figure 5, the luminescence measured immediately
after burial of this material would be about 1% of that measured had the sunlight exposure not occurred. If the dose
response is linear, it would mean that an optical dating determination would yield an age of 700 years, not zero as one
would wish. If sediment were to be buried for 1000 years, an
optical age of 1700 years would be obtained, significantly in
error. If the burial period were 50 000 years, the optical age

2. If a regeneration method is used, a test that the two dose


response curves are the same should be used. If this is not
done, or inadequately done, any age should be used with
caution.
If feldspars are measured, anomalous fading may lead to
an age that is too young. Tests can be done for the
presence of anomalous fading and, by inserting a delay of
weeks or months between irradiation and measurement,
its effect can be reduced. There is, however, no known
method of correcting for this phenomenon. If no test
results are reported, one should assume that the age may
be too young unless correct ages have been obtained from
sediments of similar material and age.
4. If feldspars are measured and the sample is relatively old
(i.e. >I00 000 years), there is evidence that the ages
obtained may be too young, perhaps because the electrons
do not remain in the traps at the environmental
temperature. Any optical ages that fall into this range
should be viewed with this possibility in mind. It is highly
desirable to have independent geochronological
information for such samples.

Sunlight exposure (s)


Figure 5. Illustration of the reduction in the 2.4 eV-excited
luminescence caused by a prior sunlight exposure. The
ordinate is the ratio of the luminescence for an aliqciot that
has had a sunlight exposure to that of an aliquot that has had
no sunlight exposure. Data are shown for quartz separated
from an approximately 500 000-year old dune sand and
Klfeldspar separatedfrom an approximately 70 000-year old
alluvial sand (based on Godfrey-Smithet al. (1988)).Note the
logarithmic axes.

D.J. Huntley and O.B. Lian

would be 50 700 years, which might not be considered significantly wrong, especially when analytical errors are taken
into account.
Alternatively, if the 70 000-year old deposit was eroded
and redeposited with 1% of the grains receiving no sunlight
exposure at all and the other 99% receiving a long sunlight
exposure, optical dating would again give an age 700 years
too large.
Clearly, then, the details of the sunlight exposure are
important. Whether or not a particular period of sunlight
exposure before burial is sufficient depends on the age of the
deposit dated and the age of the deposit from which the sediment was eroded. Ln some cases, the history of sunlight exposure and radiation dose before the last deposition will be
relevant.
There are additional considerations. The intensity and
spectrum of sunlight or daylight are highly variable, the
grains may have (or had) mineral or organic coatings on them,
or the grains may have been exposed through turbid water.
Thus, a range of possibilities of effective sunlight exposure
for the actual grains can be expected.
This leads to the question of whether or not it is possible in
the laboratory to determine if the sunlight exposure before
burial was sufficient. There are currently two possibilities.
The first follows from the fact that, within a grain, there is a
variety of traps, ranging from those which are readily emptied
by sunlight to those which are emptied with only a very long
sunlight exposure. It is this range that gives rise to the
nonexponential behaviour shown in Figure 5. During laboratory measurement, the readily emptied traps are emptied first
after the excitation is switched on, while those that are harder
to empty are emptied more slowly. The measured luminescence reflects this change, and it can be seen if an equivalent
dose versus time plot is obtained. If the equivalent dose is
constant, then all is well, but a rise in the equivalent dose
could be due to sampling of less readily emptied traps that
were not emptied before burial (as in the example above of the
70 000-year old feldspar exposed to 500 s of sunlight).
The second possibility is useful if the grains were not uniformly exposed to sunlight (i.e, some grains were well
exposed and others were not, as in the second case above). If
this has occurred, then it may give rise to data which are scattered, and can then be recognized because the deviation is
correlated with the normalization value.
The data of Huntley and Berger (1995) illustrated both
methods, but more experience has to be gained before these
techniques can be considered reliable. An alternate method is
to make measurements on a large number of individual
grains. If the relative effect of a laboratory dose shows too
wide a variation, one can conclude that the sunlight exposure
was insufficient (Lamothe, 1996).

UNEXPECTED AGES
It is inevitable that a new dating method will produce some
unexpected results. Here we describe some examples of optical ages that were very different from what was anticipated;
two are from the Palliser Triangle project. The first example
is a Holocene cliff-top dune above Onetree Creek in southern
Alberta. The optical age obtained was 112 000 15 000 years,
clearly indicating that the sediment was not exposed to sufficient sunlight during the Holocene. The source of the sediment was material that predated the last glaciation, eroded
during incision of the creek. On the basis of modern analogues it seems likely that the sediment was transported
upslope rapidly by strong winds. Since there are other examples of insufficient sunlight exposure for cliff-top dunes
(Huntley et al., 1983; Lamothe and Auclair, 1997), optical
dating should be used on such sediments with caution.

The second example is a loess from the Cypress Hills of


southeastern Alberta. Until now, this was thought to predate
the last glacial maximum (Vreeken, 1986), but our optical
ages of 12 600 1400 years for silt-sized grains and 14 000
500 years for sand-sized K-feldspar grains are leading to a
reinterpretation of the geology (W.J. Vreeken, D.J. Huntley,
and S.A. Wolfe, unpub. data, 1997).

The third example is postglacial eolian sediment (mostly


silt-sized material) from the Fraser Plateau in south-central
British Columbia, which yielded ages ranging from 22 000 to
99 000 years. Detailed investigation showed the likely reason
for the incorrect ages to be a combination of the presence of
carbonate-cemented grain clusters that were apparently
transported intact from the source material to the eolian
deposit, and a short (<I00 m) transport distance. In the laboratory, grains that had remained shielded from sunlight
within the clusters were liberated during routine sample preparation (Lian, 1997).

RECENT DEVELOPMENTS AND


PROSPECTS FOR THE FUTURE
One of the attractive features of optical dating is that ineasurable luminescence from a very small amount of sample, in
some cases as little as a single grain of sand, can be obtained.
This opens up some new opportunities. The techniques
described above involve preparing 30-60 aliquots, each holding 1000 or more of the separated grains. The use of a single
aliquot for all the measurements has been shown to be feasible for both feldspars (Mejdahl and Bgtter-Jensen, 1994;
Duller, 1995) and quartz (Murray et al., 1997). Any results
from such measurements should presently be regarded with
caution, since corrections are required because of the
repeated irradiations and heating, and it is not always clear

GSC Bulletin 534

that these were done appropriately. In addition, attention still


needs to be paid to the thermal transfer correction. Data from
a single aliquot show very little scatter, resulting in a high precision for the equivalent dose. However, this can be misleading, since scatter that results from natural variability between
grains is hidden unless several aliquots are measured. Measurements on several aliquots may also permit detection of
inadequate sunlight exposure before burial (Duller, 1995).
Extension of single aliquot techniques to measurements
on single grains may allowthe relaxation of the requirement
that all the grains were well exposed to sunlight at deposition.
As long as some of the grains were well exposed, and if it
were possible to determine which ones they were, then it
would only be necessary to determine the average equivalent
dose for these. Progress in this direction has been made by
Lamothe et al. (1994) for feldspars, and Murray and Roberts
(1997) for quartz. Measurements on single grains may also be
used to determine whether or not a sample received sufficient
sunlight exposure for dating (Lamothe and Auclair, 1997).
The time elapsed since the surface of a rock was last
exposed to sunlight is also something that, in principle, can be
determined with optical dating. This should be a fertile field
for further research if suitable material is chosen. The only
work reported so far was that of Richards (1994) on quartzite
pebbles from an archeological site in Russia.
Anomalous fading is generally observed in feldspars
(Spooner, 1992). There are two questions for which answers
are needed. The first is, "Since anomalous fading is observed
in most museum feldspar specimens, why can correct ages be
obtained for deposits from which feldspars are measured?'
The second question is, "How can one correct for anomalous
fading in the laboratory?"
Finally, a detailed understanding of the physics of the
defects, trapping, and excitation mechanisms in quartz and
feldspars is needed.

CONCLUDING REMARKS
Optical dating can now be considered a valuable technique
that can give correct ages under the right circumstances. The
researcher must be aware that the event dated is the last exposure to sufficient sunlight. It is also important to realize that
the uncertainty quoted with an age may only represent analytical uncertainties, since uncertainties in the appropriateness
of the methodology and the adequacy of sunlight exposure
prior to burial cannot be evaluated.
One should be wary of an isolated optical age with no supporting independent information. It is highly preferable that a
suite of samples, including a modern analogue (zero age) of
the source material, be measured and that there are checks
that the right ages are being obtained. Intuition or experience
is often not a good guide for deciding whether or not there was
sufficient sunlight exposure before burial for optical dating to
give the correct age.

ACKNOWLEDGMENTS
Collaboration with many colleagues over the years has led to
the development of our techniques and understanding, and
we thank them all. Financial support by the Natural Sciences
and Engineering Research Council of Canada (NSERC) and
Natural Resources Canada is gratefully acknowledged.
O.B.L. wishes to thank S.R. Hicock for his support and
encouragement. Samples mentioned here were provided by,
or obtained with the assistance of, B.F. Atwater, I. Campbell,
J.J. Clague, S.R. Dallimore, S.R. Hicock, J.T. Hutton,
E.C. Little, D.G. McPhee, J.R. Prescott, W.J. Vreeken, and
S.A. Wolfe. We thank G.R. Brooks, D.S. Lemmen,
R. McNeely, J.R. Prescott, and an anonymous reviewer for
their helpful comments on the paper.

REFERENCES
Aitken, M.J.
1992: Optical dating; Quaternary Science Reviews, v. 11, p. 127-131.
1994: Optical dating: a non-specialist review; Quaternary Science
Reviews, v. 13, P. 503-508.
Baril, M.R.
1997: O~ticaldating of tsunami de~osits:M.Sc. thesis..Simon Fraser University, ~ u r n a bBritish
~,
~dlumbia,122 p.
Berger, G.W.
1988: Dating Quaternary events by luminescence; Geological Society of
America, Special Paper 227, p. 13-50.
1995: Progress in luminescence dating methods for Quaternary sediments; in Dating Methods for Quaternary Deposits, (ed.)
N.W. Rutter and N.R. Catto; Geological Association of Canada,
Geotext 2, p. 81-104.
Better-Jensen, L. and Duller, G.A.T.
1992: A new system for measuring optically stimulated luminescence
from quartz samples;Nuclear Tracks and Radiation Measurements,
v. 20, p. 549-553.
Clague, J.J., Evans, S.G., Rampton, V.N., and Woodsworth, G.J.
1995: Improved age estimates for the White River and Bridge River
tephras, western Canada; Canadian Journal of Earth Sciences, v. 32,
p. 1172-1179.
Duller, G.A.T.
1995: Luminescence dating using single aliquots: methods and applications; Radiation Measurements, v. 24, p. 217-226.
Godfrey-Smith, D.I., Huntley, D.J., and Chen, W.-H.
1988: Optical dating studies of quartz and feldspar sediment extracts;
Quaternary Science Reviews, v. 7, p. 373-380.
Huntley, D.J.
1997: Optical dating studies of samples from the TAGLU core: preliminary results; in The Mackenzie Delta Borehole Project, (ed.)
S.R. Dallimore and J.V. Matthews, Jr.; Environmental Studies
Research Funds, Report 135, 1 CD-ROM.
Huntley, D.J. and Berger, G.W.
1995: Scatter in luminescence data for optical dating - some models;
Ancient TL, v. 13, p. 5-9.
Huntley, D.J. and Clague, J.J.
1996: Optical dating of tsunami-laid sands; Quaternary Research, v. 46,
p. 127-140.
Huntley, D.J., Berger, G.W., Divigalpitiya, W.M.R., and Brown, T.A.
1983: Thermoluminescence dating of sediments; PACT (Journal of the
European Study Group on Physical, Chemical and Mathematical
Techniques Applied to Archaeology), v. 9, p. 607-618.
Huntley, D.J., Godfrey-Smith, D.I., and Thewalt, M.L.W.
1985: Opticaldatingof sediments;Nature(London), v. 31 3, p. 105-107.
Huntley, D.J., Hutton, J.T., and Prescott, J.R.
1993: Optical dating using inclusions within quartz grains; Geology,
v. 21, p. 1087-1090.

D.J. Huntley and O.B. Lian


Huntley, D.J., Short, M.A., and Dunphy, K.
1996: Deep traps in quartz and their use for optical dating; Canadian
Journal of Physics, v. 74, p. 81-91.
Lamothe, M.
1996: Datation par les mCthodes de luminescence des feldspaths des
milieux sedimentaires:le problkme de la remise i zero; Gkographie
physique et Quaternaire, vol. 50, p. 365-376.
Lamothe, M. and Auclair, M.
1997: Assessing the datability of young sediments by IRSL using an
intrinsic laboratory protocol; Radiation Measurements, v. 27,
p. 107-117.
Lamothe, M., Balescu, S., and Auclair, M.
1994: Natural LRSL intensities and apparent lununescence ages of single
feldspar grains extracted from partially bleached sediments;
Radiation Measurements, v. 23, p. 555-561.
Leonard, E.M.
1995: A varve-based calibration of the Bridge River tephra fall; Canadian
Journal of Earth Sciences, v. 32, p. 2098-21 02.
Lian, O.B.
1997: Quaternary geology of the Fraser Valley area, Big Bar Creek to
Pavilion, south-central British Columbia; Ph.D. thesis, University
of Western Ontario, London, Ontario, 489 p.
Lian, O.B., Hu, J., Huntley, D.J., and Hicock, S.R.
1995: Optical dating studies of Quaternary organic-rich sediments from
southwestern British Columbia and northwestern Washington
State; Canadian Journal of Earth Sciences, v. 32, p. 1194-1207.
Markey, B.G., Bgtter-Jensen, L., and Duller, G.A.T.
1997: A new flexible system for measuring thermally and optically stimulated luminescence; Radiation Measurements, v. 27, p. 83-89.
Mejdahl, V. and B0tter-Jensen, L.
1994: Luminescencedating of archaeological materials using a new technique based on single aliquot measurements; Quaternary Science
Reviews, v. 13, p. 551-554.
Murray A.S. and Roberts, R.G.
1997: Determining the burial time of single grains of quartz using optically stimulated luminescence;Earth and Planetary Science Letters,
v. 152, p. 163-180.
Murray, AS., Roberts, R.G., and Wintle, A.G.
1997: Equivalent dose measurement using a single aliquot of quartz;
Radiation Measurements, v. 27, p. 171-184.

Poolton, N.R.J., Bbtter-Jensen, L., Wintle, A.G., Jakobsen, J.,


Jgrgensen, F., and Knudsen, K.L.
1994: A portable system for the measurement of sediment OSL in the
field; Radiation Measurements, v. 23, p. 529-532.
Prescott, J.R. and Hutton, J.T.
1995: Environmental dose rates and radioactive disequilibrium from
some Australian luminescence dating sites; Quaternary Science
Reviews, v. 14, p. 439-448.
Questiaux, D.G.
1991: Optical dating of loess: comparisons between different grain size
fractions for infrared and green excitation wavelengths; Nuclear
Tracks and Radiation Measurements, v. 18, p. 133-139.
Rees-Jones, J. and Tite, M.S.
1997: Optical dating results for British archaeological sediments;
Archaeometry, v. 39, p. 177-187.
Richards, M.P.
1994: Luminescence dating of quartzite from the Diring Yuriakh site;
M.A. thesis, Simon Fraser University, Burnaby, British Columbia,
96 p.
Roberts, R.G., Spooner, N.A., and Questiaux, D.G.
1994: Optical dating of quartz: a comment on Stokes and Gaylord (1993);
Quaternary Research, v. 42, p. 108-109.
Smith, B.W., Rhodes, E.J., Stokes, S., Spooner, N.A., and Aitkeu, M.J.
1990: Optical dating of sediments: initial quartz results from Oxford;
Archaeometry, v. 32, p. 19-31.
Spooner, N.A.
1992: Optical dating: preliminary resultson theanomalous fading of luminescence from feldspars; Quaternary Science Reviews, v. 11,
p. 139-145.
Stokes, S. and Gaylord, D..R.
1993: Optical dating of Holocene dune sands in the Ferris dune field,
Wyoming; Quaternary Research, v. 39, p. 274-281.
1994: Reply to Roberts et al. (1994) Optical dating of quartz: a comment
on Stokes and Gaylord (1993); Quaternary Research, v. 42,
p. 110-111.
Vreeken, W.J.
1986: Quaternary events in the Elkwater Lake area of southeastern
Alberta; Canadian Journal of Earth Sciences, v. 23, p. 2024-2038.
Wolfe, S.A., Huntley, D.J., and Ollerhead, J.
1995: Recent and late Holocene sand dune activity in southwestern
Saskatchewan; in Current Research 1995-B; Geological Survey of
Canada, p. 131-140.

GSC Bulletin 534

APPENDIX A
Terminology used in optical dating
The literature is rife with synonyms, acronyms, and specialized phrases. The following is a list of the more
common terms and their meanings:
OSL: Optically stimulated luminescence, which is the luminescence emitted when light is shone on a
sample, the emitted photons having more energy than the incident photons.
IRSL or IR-OSL: Optically stimulated luminescence with the incident light being -1.4 eV photons, which
are infrared (IR).
GLSL or G-OSL: Optically stimulated luminescence with the incident light being -2.4 eV photons, which
are green.
Anomalous fading: The phenomenon in which the measured luminescence decreases with time after irradiation and heating, even though simple theoretical calculations predict that it should not.
Bleach: A light exposure, either from sunlight or a laboratory light source.
Equivalent dose: The laboratory beta or gamma radiation dose that leads to the same amount of measured
luminescence as does the environmental dose.
Paleodose: Often incorrectly used instead of equivalent dose; paleodose equivalent would be appropriate.
Preheat: The heating of a set of aliquots before the final measurement.
Recuperation: Increase of measured optically stimulated luminescence with time or heating after
bleaching.
Saturation: The phenomenon whereby, at high enough radiation doses, further irradiation does not lead to
an increase in luminescence resulting from optical excitation.
Shine down curve: Plot of luminiscence intensity versus time, as in Figure 1B.
Stimulation: An alternative term for excitation.
Thermal transfer: Recuperation that results from laboratory heating and that would not occur in nature
during burial.
Zeroing: The phenomenon whereby exposure to sunlight reduces the subsequently measured luminescence
resulting from optical excitation. The luminescence is zero if the sample has been truly 'zeroed'.

Activity cycle of parabolic dunes based on


morphology and chronology from Seward
sand hills, Saskatchewan
Peter P. avid', Stephen A. wolfe2, David J. ~ u n t l e y ~ ,
and Donald S. ~ e r n r n e n ~
David, P. P., Wove, S.A., Huntley, D. J., and Lemmen, D.S., 1999:Activity cycle ofparabolic dunes
based on morphology and chronology from Seward sand hills, Saskatchewan; in Holocene
Climate and Environmental Change in the Palliser Triangle: A Geoscientific Context for
Evaluating the Impacts of Climate Change on the Southern Canadian Prairies, (ed.)D.S. Lernrnen
and R.E. Vance; Geological Survey of Canada, Bulletin 534, p. 223-238.

Abstract: Morphological and chronological data are used to develop a conceptual model of parabolic
sand-dune reactivation and stabilization in response to changing climate, referred to as an activity cycle.
The duration of an activity cycle is controlled by moisture availability. Optical ages from the back ridges
and dune-track ridges of adjacent dunes in Seward sand hills demonstrate that the last cycle occurred during
the nineteenth century. Ages of ridges behind the smaller of two dunes appear congruous, becoming older
with depth and younger downwind, whereas those from the larger dune record depositional events subsequent to the formation of major morphological features. Initial rates of dune advance were rapid in response
to climatic stress accumulated in the dunes. Water table fluctuations interrupted dune migration at least four
times, producing dune-trackridges. Rates of advance following formation of the first dune-track ridge averaged about 2.2 m.a-', similar to those of presently active dunes in the region.

RCsumC : Des donnCes morphologiques et chronologiques sont utilisCes afin d'Claborer un modble
conceptuel de reactivation et de stabilisation des dunes de sable paraboliques en rkaction h des changements
climatiques, processus qu'on nomme <<cycled'activitk,,. La durCe d'un cycle d'activit6 est contr61Ce par la
disponibilitC de I'humiditC. Les ages optiques des cr2tes arrigres et des crQteslatkrales de dunes adjacentes
dans les dunes Seward Sand Hills montrent que le dernier cycle a eu lieu au cours du 19esibcle. Les Bges des
crCtes situCes derribre la plus petite de deux dunes paraissent concordants, s'accroissant avec la profondeur
et dirninuant dans la direction du vent, alors que ceux de la dune plus grande tCmoignent d'Cpisodes de dCpBt
qui sont postBrieurs i la formation des principaux traits morphologiques. Les taux initiaux d'avancCe des
dunes sont ClevCs en rkaction aux contraintes climatiques accumulCes dans les dunes. Les fluctuations de la
nappe phrCatique ont interrompu la migration des dunes au moins quatre fois, engendrant des cr6tes
IatCrales. Les taux d'avancCe B la suite de la formation de la premibre cr6te latCrale sont en moyenne de
2,2 m.a-l; ils sont semblables i ceux des dunes vives actuelles de la rCgion.

' Ddpartement de geologic, Universite de Montreal, C.P. 6128, succursale centre-ville, Montreal, Quebec H3C 357
Terrain Sciences Division, Geological Survey of Canada, 601 Booth Street, Ottawa, Ontario KIA OE8
Department of Physics, Simon Fraser University, Burnaby, British Columbia V5A 1S6
Terrain Sciences Division, Geological Survey of Canada, 3303-33rd Street NW, Calgary, Alberta T2L 2A7

GSC Bulletin 534

INTRODUCTION
Recent studies have demonstrated that many sand dunes on
the Great Plains are of Holocene and even late Holocene age,
rather than being relict features formed under full-glacial or
deglacial conditions (David, 1971; Madole, 1994; Holliday,
1995; Forman et al., 1995; Muhs and Holliday, 1995;
Arbogast, 1996; Muhs et al., 1996, 1997a, b). These studies
also highlight the value of eolian deposits for paleoclimatic
reconstructions, with paleosols recording stability during
more humid periods and the sands themselves documenting
activity in response to climate or other disturbances. They
indicate that dune stabilization and reactivation occurred
repeatedly through the Holocene.
Despite our knowledge of past dune activity, our understanding of the evolution of parabolic dunes is still limited.
The morphological elements associated with modern stabilized sand dunes are the product of numerous cycles of erosion and deposition, a palimpsest of forms and features
(David, in Lernmen et al., 1998).Interpreting these features in
the context of past environmental changes requires knowledge of the processes and timing of their formation and subsequent modifications. Although eolian processes and dune
morphology in the Palliser Triangle have been well documented (David, in Lernmen et al., 1998), until recently there
were few chronological data relating to past dune activity in
this region (Wolfe et al., 1995). The paucity of organic matter
in eolian deposits means that it is rare to find suitable material
for radiocarbon dating. Where organic matter occurs in situ
within buried soil horizons, the dates obtained are for periods
of dune stability and therefore the periods of dune activity are
only broadly bracketed.
optical dating determines the time when a sediment was
last exposed to sunlight (Huntley and Lian, 1999). The technique is being increasingly applied to eolian sediments in

order to document periods of past dune activity (Prescott and


Robertson, 1997). Optical dating studies in the Palliser
Triangle were first undertaken by Wolfe et al. (1995), who
suggested that the most recent period of widespread dune
activity occurred within the last 200 years, with previous
activity occurring in the early to middle part of the last
millennium.
The purpose of this paper is to present a conceptual model
of parabolic dune reactivation and stabilization in response to
changing climate, herein referred to as an activity cycle. The
model is based on the morphology and chronology of back
ridges and dune-track ridges, which are indicators of spatial
and temporal fluctuations in groundwater levels and, consequently, past climatic events.

STUDY AREA
Seward sand hills, located in southwestern Saskatchewan, are
the southeasternmost major dune occurrence within the
Palliser Triangle (Fig. 1). Extending northeast from near the
abandoned town of Antelope, the sand hills form an elongate
body about 30 krn long and 3-5.5 km wide (Fig. 2), lying
almost parallel to the north slope of the Swift Current Creek
Upland, situated 5-8 krn to the south. The orientation of the
dune area reflects both availability of deflatable material and
prolonged sand transport by southwesterly winds.

Physiography and Quaternary geology


The sand hills lie within the Bigstick Lake Plain physiographic subdivision of Saskatchewan (Acton et al., 1960).
Low-lying flat surfaces dominate, although till knobs and low

Figure 1. Location of Seward sand hills and other major sand dune occurrences
(shaded) in the Brown Chernozemic Soil Zone. Inset map relates this soil zone to
margins of Palliser Triangle (heavy dashed line). Star on inset is approximate
location of study site.

P.P. David et al.

mounds of ice-contact glaciofluvial deposits commonly


extend above the relatively thin eolian sediments. Remnants
of collapsed ice-contact channels are also evident.
The eolian sand deposits are generally fine to medium
grained. In the western third of the area, they overlie
glaciolacustrine sand and silt, whereas, in the east, they overlie mostly fine-grained glaciofluvial deposits (fine sand to
fine gravel) and local occurrences of glaciolacustrine sand
(Fig. 2). These source deposits provide a finite supply of sand
for deflation and dune building (cf. Muhs and Wolfe, 1999).
Deflation of the finer material has produced pebbly and,
locally, cobble-sized lag concentrates. Inasmuch as most of
the region has been affected by wind since deglaciation, a
veneer of eolian sediment is found almost everywhere on the
surrounding areas.

of annual precipitation to potential evapotranspiration is


about 0.70, it can drop below 0.50 during drought years such
as 1988, resulting in short-term semiarid conditions (Wolfe,
1997). Under the prevailing climate, active sand dunes
remain moist even though the vegetated surfaces may be subject to water deficiency (cf. Prill, 1968; Winter, 1986).
The wind-energy regime of the area is relatively high,
with strong winds occurring in spring and fall as a result of
pressure gradients, and from Chinooks blowing eastward
from the Rocky Mountains (Wheaton and Chakravarti,
1987). The dominant dune-forming winds (>5.8 m.s-') at
Swift Current are from the southwest most of the year, but are
generally more southerly in early spring.

Drainage and hydrogeology


Bridge Creek traverses the western extreme of the dune area
in an incised meandering channel, flowing north into
Antelope Lake (Fig. 2). Tenrnile creek, a small ephemeral
stream, flows north into the flat, central part of the sand hills
via a shallow entrenched channel which is ponded into small
temporary lakes connected by discontinuous channel fragments among the sand dunes. Farther north, the entrenched
channel reappears and exits the dune area (Fig. 2), suggesting

Climate
Swift Current, 30 km east of Seward sand hills, receives
approximately 375 mm of precipitation annually, about
one-third of which falls as snow (Environment Canada,
1993). Mean annual potential evapotranspiration is about
546 mm, resulting in a subhurnid to dry subhumid moisture
regime (Wolfe and Lernmen, 1999). Although the mean ratio

gure :

108"20'

T
I

Dunesand
Loess
Sandur deposits
(sand and gravel)
Glaciofluvial deposits

108"IO'

fl

Lacustrine sand
wlth mlnor s ~ l t
Thick lacustrine silt

Mean dune
orientation

Alluvium

Escarpment

Y-U-

Eroded slope

Road

Figure 2. Surficial materials of Seward sand hills and surrounding areas (after
David,1964).

GSC Bulletin 534

that the channel predates dune formation. The ephemeral


lakes were formed by dunes migrating into the channel, a
comlnon phenomenon in sand hill areas (Dobrovolny and
Townsend, 1947; Molnhr, 1961; Lanner 1992; Loope et al.,
1995).
nature of
lakes in the sand
The
reflects variability in the position of the water table. High
phreatic water levels flood interdune areas and blowout
depressions, whereas lowering of the water table under conditions of drought results in drying of the lakes. This association is well evidenced by aerial photographs. The earliest
available photographs, taken following the 1930s drought,
show all lake basins and blowout depressions dry. In contrast,
photographs from the mid-1950s, following a comparatively
humid period, show standing water in all depressions and
lake basins. Groundwater is recharged by both precipitation
and stream flow. Recharging by precipitation is more efficient through active dune sand (cf. Gupta, 1979; Berger,
1992), especially from slow melting of snow accumulated in
the lee of the slip face. Once dunes become stabilized, the distribution of surficial recharge water is controlled by vegetation. Transpiration often prevents recharge water from
reaching the saturated zone, so that extended drought may
result in the complete exhaustion of near-surface groundwater under a vegetated dune (Gupta, 1979).

EOLIAN LANDFORMS
Seward sand hills comprise a variety of eolian landforms.
Erosional features predominate in the southwest and include
blowout hollows, pits, and elongate depressions bordered by

Unfilled

1.
':

MS,
. .,J

blowout ridges (David in Lemmen et al., 1998). In the central


area, well developed parabolic dunes are the most common
features. Most of the dunes are 'unfilled', while a few are
'partially filled' with sand (inset, Fig. 3; Wolfe and David,
1997; David in Lemmen et al., 1998), indicating that sediment supply in the area is moderate to low (cf.-wolfe and
David, 1997). The northeastern extension of the sand hills
forms a dune tongue, a colnmon feature downwind of dune
occurrences (David, 1993, in Lemmen et al., 1998).
Most of the dunes in Seward sand hills are presently stabilized, with active sand restricted to blowouts and active heads
of otherwise vegetated dunes (light tones in Fig. 3). This
blowout activity causes disintegration of the well developed
parabolic dunes, resulting in a more chaotic landscape with
numerous dune remnants and blowout features (cf. Wolfe and
David, 1997). There has been a net trend toward dune stabilization over the past 70 years, both locally and in the Great
Sand Hills region (Fig. 1; Wolfe et al., 1995). These observations suggest that the present level of eolian activity cannot
account for the formation of the well developed stabilized
dunes.
Although dune axes are oriented northeast, more or less
parallel to the elongation of the dune field (Fig. 2), a large
number of the blowouts trend southeastward. In addition, a
few stabilized dunes are slightly asymmetric to the southeast,
signifying deformation in that direction prior to stabilization.
Finally, most of the active dunes in the area show evidence of
secondary sand transport to the southeast, transverse to dune
axes. The northwesterly deforming winds are probably
short-lived deviations from the principal southwesterly

Aerial photograph of the central, low area of


Seward sand hills showing the seven sand dunes
with prominent back ridges and dune-track
ridges. Light tones denote areas of active sand;
darkest tones denote small ponds and other wet
ground. Locations of features sampled for optical dating are indicated by arrows (dunes 1 and
2 ) . Abbreviations: BR, back ridge; DT,
dune-track ridge; BS, back slope; MS, modern
(control) sample. Excavation pits were located
along the central axis of each dune. Inset presents a schematic diagram of parabolic dunes in
plan view, following classification of Wolje and
David (1997) and David (in Lemmen et al.,
1998). NAPL A21004-63 (taken in 1969).

P.P. David et al.

dune-moving winds, because small blowouts take a relatively


short time to develop and the presently active sand areas can
be readily deformed by cross winds.
Microenvironmental differences dictate that even adjacent dunes will not be exactly alike. Even within an individual
dune, various microenvironments cause spatial variability in
the recurrence, duration, and intensity of eolian processes.
Local dependent variables include rnicrotopography, surface
slope and aspect, sand thickness, association with noneolian
sediments, subdune topography, moisture content of the dune
and subdune sediments, depth to water table, type and density
of vegetation cover, and animal and human activity. The consequent differences in the time, duration, and intensity of
exposure of the sand to sunlight may result in unexpected, but
correct, optical age determinations (Huntley and Lian, 1999).

Back ridges
Back ridges are low, arcuate sand accumulations, concave
downwind, that connect the wings of parabolic dunes around
the upwind margin of blowout depressions (Fig. 3; David in
Lernmen et al., 1998). First recognized as morphological elements of parabolic dunes by David (1964), they may form
either through reactivation of a stabilized dune or as part of a
newly formed blowout dune (Jennings, 1965; Pye, 1980;
Fryberger et al., 1984). The f i s t scenario indicates a significant, long-term climate shift, while the second scenario may
reflect only local disturbance that need not be climatically
controlled. In both cases, the ridge develops under the influence of varied wind directions (David, 1988). Vegetation on
the ridge traps eolian sand, building the ridge vertically and
laterally. The significance of a back ridge as a morphological
time marker is first recognized in this paper in light of the
available age data.

Dune-track ridges
Dune-track ridges are arcuate, sometimes slightly sinuous or
irregular ridges connecting the two wings across the blowout
depression (Fig. 3; David in Lernmen et al., 1998). They mark
the former position and shape of the back base line of the dune
head, first recognized by Kerr and Nigra (1952) for arid
(desert) dunes and by David (1964) for parabolic dunes.
These ridges tend to be either concave downwind or straight,
and rarely convex. The crest lines of dune-track ridges are
generally even, contrasting with the irregular crest line of
back ridges. The plan and profile shapes reflect the lower
back of a parabolic dune under conditions of abundant sediment supply, with the upwind half of the dune resembling that
of a barchan in plan view.
Dune-track ridges develop when a period of dune rnigration is interrupted by more humid intervals that allow the
water table to rise near or above the ground surface. Increased
moisture availability promotes rapid development of vegetation on the basal portion of the dune. Although the rest of the
dune surface remains barren, activity will be reduced by the
greater interstitial moisture in the sand. A subsequent return
to more arid conditions will reinitiate more rapid dune

migration, leaving behind a low ridge protected by vegetation. This vegetation is able to survive the transition to more
arid conditions owing to the strong lateral component of
groundwater percolation in active dunes (McCord and
Stephens, 1987). Furthermore, the root system of the vegetation may extend into the finer grained subdune deposits
where more moisture is available (Winter, 1986). Once a
dune-track ridge is established, the vegetation on it may trap
new sand blown onto it, building it up and further protecting it
from erosion.
Dune-track ridges do not reflect a substantial change in
climate, since only a small basal portion of the dune becomes
stabilized. Rather, their development marks short-term climate fluctuations, since a long-term increase in moisture
availability would cause the entire dune to become stabilized.
The size and number of dune-track ridges that develop
depend on the size of the dune and its height above the fluctuating groundwater level. A number of small ridges may form
behind a large dune if it is located close to the water table but
remains active during more humid intervals. A smaller dune
under the same climatic setting may be stabilized completely.
The formation of dune-track ridges should immediately
postdate an interval of more humid climate because development and preservation of these ridges require relatively dense
vegetation growth, which in turn is controlled by the phreatic
groundwater level. As the ridges are erosional features and
may be modified by subsequent deposition, the sand which
composes them may relate to a number of depositional
events.

METHODS
A set of seven morphologically similar dunes in the central
portion of Seward sand hills was selected for study (Fig. 3).
Each of the selected dunes features a well developed back
ridge and a set of four equally well formed dune-track ridges.
All of the dunes are mostly or completely stabilized, with the
exception of dune 2, which has a large active surface on the
head. The active surfaces of dunes 1 and 3 that are evident in
Figure 3 became mostly stabilized since the aerial photograph
was taken in 1969.

Field
Distances between ridges along the axes of dunes 1 and 2
(Fig. 3) were measured in the field, as were ridge heights and
slope angles. Distances at the other five dunes were measured
photogrammetrically using a WildTMstereometer. These values were converted to ground distances based upon field measurements between fixed points readily identifiable on the
aerial photographs, yielding a field precision of better than
1 m. Ridge lengths (arc distance) were measured photogrammetrically for dunes 1 and 2, although the length of the back
ridge of dune 2 is difficult to ascertain because it is partially
covered by dune 1 in the south and by another dune in the
north (Fig. 3). Similar uncertainties preclude meaningful
measurement of ridge length for the other five dunes.

GSC Bulletin 534

Samples from the back ridge and dune-track ridges of


dunes 1 and 2 were collected for optical dating. Pits were
hand dug in each ridge to a depth of about 80 cm, and all visible sedimentary structures noted. Samples were collected at
25,50, and 75 cm depth for analysis of grain size and moisture
content. Next, samples for dating were collected at 50 cm
depth, under a tarpaulin to preclude light infiltration, and
placed in sealed 1 L light-tight containers. The uniform sampling depth was selected to eliminate bias in dating. However,
as initial results indicated that the ages of the shallow samples
from the back ridges of both dunes were significantly younger than those from many of the dune-track ridges that lay
downwind, two additional samples were subsequently collected at depths of 1.0 and 1.5 m from the back ridge of dune
2, to examine how age can vary with depth.

in Table 1. This procedure may have introduced a systematic


error of the order of 5%. However, when scaling of the additive dose and regeneration data were introduced as a parameter in the fit, the value was within 20 of unity for all but one of
the 14 samples (o ranging from 3 to 6%). However, a systematic deviation from unity occurred, with the average value
being 1.045 f 0.012. This is the value by which the intensities
of the regeneration data must be multiplied to yield a best fit
to the additive-dose data. The equivalent doses shown in
Table 1are those obtained with this parameter fixed atunity; a
correction has not been made because it is not certain that this
is not an artifact of the fitting procedure arising from an inappropriate assumption about the statistics. In all cases, the
equivalent doses obtained using the two methods were very
similar.

In addition to the samples from the back ridges and


dune-track ridges, one sample was collected from near the
base of the back slope of dune 2 at 50 cm depth, and another
from its presently active surface to serve as a modern
analogue.

Environmental beta and gamma dose rates were calculated from measured K, U, Th, Rb, and moisture contents
(Table 2), using the conversion factors of Nambi and Aitken
(1986) and the beta attenuation factors of Mejdahl (1979).
Internal beta dose rates from K and Rb were calculated as outlined in Ollerhead et al. (1994). Cosmic-ray dose rates were
estimated according to the prescription of Prescott and
Hutton (1988). The total dose rates so calculated are given in
Table 1.

Optical dating
For optical dating, K-feldspar grains in the 180-250 pm fraction were separated. Extracted grains were excited with
1.4 eV photons and emission of 3.1 eV photons was measured. Equivalent doses (Table l) were determined using
both the additive-dose and the regeneration methods for each
sample according to techniques developed by Ollerhead et al.
(1994) and Huntley and Clague (1996). These are described
in the review by Huntley and Lian (1999). An infrared bleach
was used in both methods and the preheat was at 120C for 16
hours. A typical example of the additive-dose data is shown in
Figure 4. The regeneration data yielded much better precision
than the additive-dose method and, since this precision is
desirable for age comparisons, the equivalent doses (De4)
obtained in this manner were used to calculate the ages shown

Optical ages were calculated by dividing the equivalent


doses by the dose rates (Table 1, optical age 1). Because several of the contributions to the uncertainties in these ages are
of a systematic nature, we have also calculated a second set of
ages, using the average calculated dose rate of 3.08 Gy.ka-',
in the expectation that these will give a more precise estimate
of the relative ages (Table 1,optical age 2). The uncertainties
quoted are only those arising from the uncertainties in the
equivalent doses; these ages should thus be used for relative
comparisons only.

Table 1. Radiation dose rates, equivalent doses, and

optical ages of the dune sands from Seward sand hills.

Locatlon

Optical age
(years before AD 1995)

Sample no.
(SAW-)

Depth
(m)

Dose ratea

94-30
94-31
94-32
94-33
94-34

0.5
0.5
0.5
0.5
0.5

3.15
3.08
3.13
3.13
3.02

0.387f0.015 123f6
94 f 6
0.288f 0.015
0.363+0.019 116f 7
94 f 6
0.295f0.015
0.460f 0.019 152 f 8

94-35
95-10
95-11
94-36
94-37
94-38
94-39
94-40

0.5
1.0
1.5
0.5
0.5
0.5
0.5
0.5

3.12
3.07
3.04
3.11
3.11
2.99
3.10
2.99

0.390f0.010
0.553 f 0.016
0.562 f 0.016
0.542-tO.018
0.523f0.015
0.477 f 0.026
0.426f0.025
0.351 +0.019

Deq
(~~.ka-~f0.10) (Gy)

zb

Dune I
BR 1
DT 2
DT 3
DT 4
DT 5

126f5
84 f 5
118f6
96f 5
149 f 6

Dune 2
BR 1
BR 1
BR 1
DT 2
DT 3
DT 4
DT 5

BS 6

125f5
180 f 8
185 f 8
174i8
168f7
160 f 10
137 f 9
117f7

127f 3
180 f 5
182 f 5
176f 6
170k5
155 f 8
138 f 8
114f 6

a Includes an estimated equivalent alpha dose rate of 0.06 f 0.03Gyka".

Relathre optical dates, calculated using the same dose rate for all. that do not include the
uncertainty in the dose rate.
Dose rate for 94-41had it been burled 50 cm. for comDarison DurDoses onlv.

Dose (Gy)
Figure 4. Typical data for an equivalent dose determination
using the additive-dose method. The data shown are for
sample SAW94-36, and are for the first 10 seconds of
excitation, during which the luminescence decreased by 20%.
Linearfits to the data are shown and were obtained assuming
maximum likelihood statistics.

P.P. David et al.

Table 2. Potassium, uranium, thorium, and water

contents of sand samples.

Table 3. Morphological data for all dunes studied.

K (% f 5%)
Sample no.
(SAW-)

94-30
94-31
94-32
94-33
94-34
94-35
95-10
95-11
94-36
94-37
94-38
94-39
94-40
94-41

Whole
sample

1.49
1.49
1.50
1.48
1.48
1.55
1.48
1.49
1.47
1.48
1.42
1.45

Surroundings

1.54
1.50
1.43
1.50
1.47
1.56
1.51
1.49
1.43
1.48
1.56
1.49

U
(pgg-')

1.34 f 0.06
1.22+ 0.06
1.36i! 0.06
1.33i 0.06
1.12i0.06
1.21 0.06
1.16*0.09
1.13f 0.09
1.28? 0.06
1.35i! 0.06
1.11?0.06
1.34f 0.06

Th
(clg.9")

3.5k 0.05

1.06f 0.09

3.1 0.1
3.9-1: 0.05

Water

Location

Individual1

0.054
0.060
0.067
0.057
0.063
0.065
0.035
0.035
0.048
0.070
0.058
0.062
0.055
0.049

'A = (mass of water)/(dry mass); uncertainties were taken to be 0.020.


K contents of the grains measured were taken to be 12.5 f I%, on the basis of K

measurementsand element mapping, uslng a scanning electron microscope, of 21


grain separates, including two from this set (Huntley and Baril, 1997).
Rb contents of the grains were taken to be 420 vg.g-l, on the assumption that the
Rb:K ratio of the grains was the same as that for the sand.

The initial justification for use of a fixed dose rate is based


on the expectation of uniformity of the sand. Sources of (and
uncertainties in) the dose rates are: K and Rb contents of the
feldspar grains (0.06 Gy.ka-l), whole sample K contents
(0.06 Gy.ka-l), water contents (0.04 ~ ~ . k a - U
l )and
, Th contents (0.03 Gy.kxl), and alpha dose rates (0.03 Gy.ka-l).
Sample-to-sample variations in the K and Rb contents of the
feldspar grains and in the alpha dose rates are probably much
smaller than our estimates of the uncertainties. The differences in the K contents shown in Table 2 are substantially less
than the analytical uncertainties, and the true variation may
be smaller still. The estimate of the uncertainty in the
long-term average water content is probably much larger than
the sample-to-sample differences. From the association
between groundwater table fluctuation and dune-track ridge
formation, it is obvious that the sands enclosed in them were
periodically either saturated or simply moister than today.
Aerial photographs demonstrate that the ridges have undergone similar periodic conditions of high groundwater table
since their formation (see 'Climate' section). The final justification for presenting ages calculated with a fixed dose rate
lies in the consistency in the sequence of ages for dune 2 and
the similarities of ages obtained from both dunes for the tops
of the back ridges and for dune-track ridge 5.
Measurements on a zero-age sample, SAW94-41, yielded
a nonzero equivalent dose of 0.052 k 0.014 Gy. It is possible,
therefore, that the other samples suffer from a zero-age error
of this order and that 0.05 Gy should be subtracted from all
the equivalent doses. We have not done this because 1) sample SAW94-41 was from the head of the active portion of
dune 2, which, although under the present conditions normally receives the most sunlight exposure, is not a true analogue of a dune-track ridge deposit; 2) another zero-age
sample from an active dune surface in the Burstall sand hills

Distance from back ridge (m)

Dune 1
BR 1
DT 2
DT 3
DT 4
DT 5

BS 6
Dune 3
BR 1
DT 2
DT 3
DT 4
DT 5
BS 6
Dune 4
BR 1
DT 2
DT 3
DT 4
DT 5
BS 6
Dune 5
BR 1
DT 2
DT 3
DT 4
DT 5
BS 6
Dune 6
BR 1
DT 2
DT 3
DT 4
DT 5
BS 6
Dune 7
BR 1

DT 4
DT 5

Cumulative

Ridge
length (m)

%2

0.0
53.3
15.1
33.2
28.5

0
53
68
102
130

0
41
53
78
100

400
297
275
207
126

52.5

210

0
83
17
32
29
90

0
83
100
132
161
251

0
52
62
82
100

0
63
18
34
30
95

0
63
81
115
145
240

0
43
56
79
100

0
61
16
37
32
97

0
61
77
114
146
243

0
51
15
33
25
128

0
51
66
99
124
252

136
175

100

0
42
53
78
100

0
41
53
80
100

Determined from distance between centre lines of adjacent ridges.


Calculated as a proportion of the distance between BR 1 and DT 5.

(Fig. 1) yielded an equivalent dose of -0.01 f 0.02 Gy; and


3) there is no obvious argument that the 'zero' error should be
the same for all samples, unless it is truly zero, since it is
dependent on the sunlight exposure prior to burial. In fact,
one day prior to sampling, strong winds may have exposed or
deposited the surface sand which then subsequently became

GSC Bulletin 534

Our optical ages are of the order of 100-200 years, a time


scale for which adequate tests for absolute accuracy of the
method have not yet been made. We can argue, however, that
the ages are credible on the basis of a near-zero age obtained
for the modern sample, and the credible ages and age comparisons to be found for much older samples in Hiitt et al. (1988),

wet, all the while not being exposed to direct sunlight due to
continuous cloudy conditions. In retrospect, either more
modern analogues should have been measured or sampling
should have been done following a period of exposure of the
sand to bright sunlight.

Calculated best-fit lines shown

25

50

75

100

125

150

175

200

225

Ridge distance (m)

40

80

120

160

200

240

Dune 1 - ridge distance (m)

Figure 5. Morphological characteristics of dunes 1 and 2 (see Fig. 3for locations): A) ridge
arc-length versus distance from back ridge; B) distancefrom the back ridges. Note the large
initial distance covered by dune 2, while dune 1 'catches up' only after the last dune-track
ridge (DT5),arriving at nearly the same final distance at the back slope (BS 6).Dotted line is
shown for reference purposes only.

P.P. David et al.

The heights of the dune-track ridges range from 0.50 to


1.35 m, whereas the corresponding slope angles range from 4
to 14", showing a greater spread in slope angles associated
with lower ridge height (Fig. 6). Such a spread in angle values
is due to the small size of the ridges, the mostly erosional origin of the ridges, and the subsequent deposition of eolian sand
onto a vegetated surface. Were the features higher and purely
constructional in origin, sediment transport processes would
have produced a back slope and a frontal slip face on which
slope angle variations are more subdued (cf. David, 1981).

Spooner et al. (1990), Aitken and Xie (1992), Clarke (1994),


Duller (1994), Fuller et al. (1994), Lian et al. (1995), Huntley
and Clague (1996), and Lian (1997).

RESULTS
Morphometry
The distances between adjacent ridges and the arc lengths of
individual ridges of dunes 1 and 2 were used to evaluate the
morphometric relationships of the two dunes (Table 3).
Best-fit lines calculated for the data of both dunes converge
toward zero near the base of the modern back slope (Fig. 5A).
The length of the back ridges is a measure of dune size, and
the difference between the two ridges indicates that the smaller
dune (dune 2) is about 70% of the size of the larger one.

Sedimentary structures
The sand in most exposures is oxidized, light brown to yellow-grey changing to brown-grey with depth. There are no
clearly visible bedding planes or laminations in any of the
exposures, except for thin layers of light, bright grey (2.5 YR
711; MunsellTM Colour Chart), unoxidized sand produced
when rain leaches away the oxidized coatings from sand
grains on the dune surface (David, 1968). In section, these
layers are mostly planar tabular, although two have a wavy,
rippled appearance and only a few pinch out within the exposure. The inclinations range from 0 to 1lo, with most being
about 6". The majority (75%) of the layers dip upwind (southwest; Fig. 7).

The smaller dune (dune 2) shows a greater distance from


the back ridge to the first dune-track ridge than dune 1,
whereas the larger dune shows a greater distance between the
final track ridge (DT 5) and the back slope (BS 6), such that
the total distance from the back ridge to the back slope of the
two dunes is very similar (205 and 210 m, respectively;
Fig. 5B). Similarly, for dunes 3 through 7, distance values
vary only from the back ridge (BR 1) to the first dune-track
ridge (DT 2), and from the last dune-track ridge to the back
slope (Table 3). Approximately 40 to 50 per cent of the distance between the back ridge (BR 1) and the last dune-track
ridge (DT 5) was covered by the dunes in their first period of
migration from the back ridge to the first dune-track ridge
(DT 2), while subsequent migration distances were smaller
with less variation among the dunes (Table 3).

Chronology
The optical ages of the back ridge and dune-track ridges of
dune 2 show a congruent series, starting at 185 k 8 years for
BR 1 (1.5 m depth) and becoming systematically younger
downwind, terminating at 117 7 years at the back slope of
the present dune (Table 1). When the second set of optical

16 I

14 -

I I

12 -

-a,

0,

10-

C
(U

a,
a
0

Figure 6.

Calculated best-fit line

8-

6I

4-

0.2

0.4

0.6

0.8

1.O

1.2

Ridge height (m)

1.4

Relationship between slope angle and ridge


height of dunes 1and2 along a central traverse.

GSC Bulletin 534

ages (Table 2) is plotted as a function of distance from the


back ridge (Fig. 8A), they show a remarkable seriation. From
the linearity of these data, we deduce that the precision of the
ages is indeed as calculated using a fixed dose rate. In contrast, ages for the same features of dune 1 do not show a coherent pattern. Although samples from 50 cm depth in both back
ridges yielded the same age, as did those from the last
dune-track ridge (DT 5), samples from the other three
dune-track ridges did not yield comparable age pairs as
anticipated. The 1.0 and 1.5 m deep samples from the back
ridge of dune 2 yielded overlapping ages, significantly older
than that for the 0.5 m deep sample (Table 1).

The optical ages all record eolian activity within the last
two centuries. The best estimate of the start of this activity is
185 8 years ago (ca. AD 1810), as recorded by dated samples from the back ridge of dune 2. It is unlikely that the climatic event that triggered this activity happened at that time,
but rather sometime earlier, with a response time of perhaps
10-15 years necessary to activate formerly stabilized dunes
(discussed in more detail later). Although morphological evidence strongly suggests synchronous development of the
dunes, the optical age sequences obtained from the two dunes
are different. The fact that the ages from dune 2 become consistently younger downwind suggests that they reflect ridge
formation, and hence may date climatic events. However,
with the notable exception of DT 5, the ages obtained from
the dune-track ridges of dune 1 are not the same as for corresponding ridges of dune 2, nor do they show the expected pattern of being younger downwind. Nonetheless, all the dates
are valid in the sense that each reflects the time of last exposure of sand to sunlight (Huntley and Lian, 1999), and consequently do represent eolian events. We believe that the
differences between the two dunes suggest that ages from the
smaller dune (dune 2) relate to the climatic shifts that caused

INTERPRETATION
The seven dunes examined in this study occur within a relatively small area and therefore must have been affected by the
same climatic events. If the major morphological features of
the dunes are the product of climate-controlled processes, as
discussed previously, then the striking similarities of the
dunes suggest that they evolved synchronously and that their
constituent elements formed at roughly the same time. Most
noteworthy are the well developed back ridges and
dune-track ridges. For each dune, the back ridge, with the
more irregular profile, formed at the upwind margin of the
blowout depression. The back ridges are interpreted to represent the last position of the upwind edge of stabilized dunes
prior to reactivation and therefore are expected to be the oldest morphological elements of the dunes. This is supported by
optical ages of 180 f 8 and 185 f 8 years from 1.0 and 1.5 m
depths in dune 2, respectively. Although the two ages cannot
be differentiated at one standard deviation, the law of

Dune 1
BRI

superposition dictates that the upper sand is younger, a conclusion that the optical ages do not contradict. The dune-track
ridges are smaller and more uniform in cross-sectional profile
than their associated back ridges. The dominance of
low-angle, southwest- (upwind) dipping grey layers within
ridge sediments suggests that they accumulated on the stoss
side of the dune near the base of the back slope and, as such,
could not be former frontal deposits (cf. McKee et al., 1971;
Hunter, 1985; Halsey et al., 1990).Their preservation reflects
stabilization of the lower back slope of the dune by vegetation
under conditions of temporarily high water levels.

Dune 2
DT2

DT3

DT4

DT5

BR1

DT2

DT3

DT4

DT5

SWDune sand with

grey layers;
inclination value
to SW or NE

5"

Sample location

123 5

Optical age (years)

Figure 7. Stratigraphy of excavation pits and optical ages for dunes 1 and 2. See Figure 3for site
locations.

P.P. David et al.

formation of the dune-track ridges, whereas those from the


larger dune (dune 1) date minor depositional events that postdate initial ridge formation. Therefore, the age differences
reflect microenvironmental differences between the two
dunes, a common phenomenon in nature (cf. Phillips et al.,
1996).

A'1 of the ridges show mor~ho'ogical evidence of


modification following their initial formation, as supported
evidenceshowing the large 'pread of 'lope
versus ridge height (Fig. 6; cf. David, 1981), and by aerial photographs. In the case of back ridges, the irregular, sinuous base

lines and crest lines indicate that sand was either removed
from the ridges by erosion, including rainwash, or blown onto
them at random. These processes may have occurred repeatedly during the many renewed periods of dune activity
(David, 1993), but would have been most significant when
vegetation cover was sparse. The dune-track ridges were less
affected by postform~tional modification. ~ i r i a lphotographs taken in 1939 show active sand along the top of some
of these ridges, indicating that they either received sand or
had undergone deflation during the preceding drought interof stratification or lamination in the
v a l An apparent
looser near-surface sand (Fig. 7 ) is probably due to slow

50
100
150
Distance from back ridge (m)

200

Distance from back ridge (m)

Figure 8. Optical age 2 (Table 1) versus distancefrom back ridge for the dune-track
ridges and back slope: A) Dune 2 - the ages are based on the assumption of the same
radiation dose ratefor all samples and do not include uncertainties in the dose rates.
The zero error in the optical ages is unknown, but may be as much as 20 years. The
slope of the best-fit linear regression line (dashed, R-squared 0.99) is 2.2 f 0.2 m.a-l,
this being the average migration rate of the dunefrom DT2 to BS 6. Note that the back
ridge sample lies to the left of this line, indicating that the initial migration rate was
more rapid. B) All seven dunes (based on the assumption of synchronous evolution) Dune 2 is the bold dashed line. Lines connect mean ages (arrows);standard errors are
the same as depicted in (A).Note similarities between dunes, with the most significant
differences occurring aferformation of DT 5. This may indicate thatformation of the
back slopes was not synchronous.

GSC Bulletin 534

deposition onto vegetation. Postformational changes are


apparently influenced by ridge height (Fig. 6). Low ridges are
generally less vulnerable to wind erosion because they
receive greater protection from low-growing shrubs, whereas
higher ridges commonly have only a grass cover, which is
more easily destroyed. Since back ridges are generally larger
than dune-track ridges, they appear to be subject to greater
modification, resulting in a more complex micromorphology.

that may be synchronous across the study area, the timing of


final back slope stabilization is a function of sediment supply
related to the rate of spread of vegetation encroaching on the
dune surface. Local variability in stabilization is evidenced
by the ongoing blowout processes that affect some dunes to
varying degrees (e.g. dunes 1,2,3,5, and 6, Fig. 3). while others are stable (e.g. dunes 4 and 7, Fig. 3). As a result, the age of
the back slope is unlikely to be the same for each of the dunes
studied.

Dune migration

DISCUSSION

The average migration rate of dune 2 during the last period of


major advance is estimated to have been between 2.5 and
4.0 m.a-l, based upon the ages of the back ridge (185 f 8
years) and the back slope (117 f 7 years) of the dune. These
values are similar to rates of modern dune advance in the
Palliser Triangle, measured over periods of greater than
30 years (David, 1964, 1972; St-Onge, 1966; Wolfe and
Lernmen, 1999).

To place the morphological characteristics of dunes in the


Seward sand hills, and parabolic dunes in general, within the
context of landform evolution, the concept of a dune activity
cycle is introduced. An activity cycle is defined as the succession of eolian events between two successive phases of complete dune stabilization, and is illustrated in Figure 9.

Average migration rates between successive control


points are not possible to calculate, because in most instances
the optical dating analyses were incapable of differentiating
the ages of these features (the optical ages overlap at one
standard deviation; Table 1, Fig. 8A). Nonetheless, plotting
the age of the samples versus distance from the back ridge
indicates that the migration rate of dune 2 averaged about
2.2 m.a-l between DT 2 and BS 6 (Fig. 8A). The fact that the
sample from BR 1 lies well to the left of the best-fit regression
line for samples from DT 2 to BS 6 suggests that the initial
migration rate (between BR 1 and DT 2) was more rapid,
likely reflecting the strongly negative groundwater budget
that resulted from a severe drought preceding dune reactivation. A minimum estimate of this rate is roughly 4 m.a-l,
using the two standard deviation age difference (Table 1,
optical age 2). Additionally, the fact that disruption of the stabilizing vegetation cover would have exposed a relatively
thick sequence of dry sand to the wind suggests that transport
may have occurred at rates similar to those observed in arid
regions (approx. 5-8 rn-a-l;Finkel (1959), Hastenrath
(1967), Bordwka (1979), Haynes (1989)), where activity is
transport limited and not impeded by moist sand. A decrease
in dune migration rate following formation of the first
dune-track ridge (Fig. 8) relates to the first period of migration after moisture had re-entered the surface of the dunes.
Under such conditions, sand movement would become
increasingly supply limited, with humidity becoming a more
dominant factor than wind in controlling sand transport (e.g.
Muhs and Wolfe, 1999).

Stabilized sand dunes will not immediately become active


in response to an increase in aridity. Rather, any activity will
be delayed owing to protection against deflation provided by
living plants, dead rooted vegetation, and dead vegetation
debris on the stabilized dune surface (Fig. 9, stages 1 and 2).
Although sand within a stabilized dune loses much more
moisture through transpiration, and receives and retains less
moisture from precipitation, than does exposed sand (Loewy,
1953; Prill, 1968; Gupta, 1979; Keen 1987; McCord and
Stephens, 1987; Keen and Rosenberry, 1988; Berger, 1992),
it is still more difficult to initiate activity on stabilized dunes
than to cause the expansion of an already active sand surface.
However, with increasing duration of drought, vegetation
will gradually exhaust the available moisture of the dune
sands to the base of the roots. The depth of moisture exhaustion, or the resulting thickness of dry sand, may be viewed as a
measure of the climate stress accumulated in the dune. A triggering action such as wind shear, fire, animal or human activity, or a combination of these may then disrupt the vegetation
cover. Under the present climate regime, neither fire alone
(David, 1969) nor short-term drought like that of the 1930s
and 1980s is likely to start activity on undisturbed stabilized
parabolic dunes. However, once activity has started, it
quickly extends over the entire dune because the sand below
the surface has become dry, and therefore conditioned for
easy transport (Fig. 9, stage 3). The initial mobilization of the
sand results in relatively high migration rates during the early
stages of dune advance (Fig. 8), and in the formation of the
back ridges at the upwind edge of the blowout depressions.

If synchronous evolution is assumed, the relationship


between optical age and distance from BR 1to DT 5 is similar
for all seven dunes (Fig. 8B), with dune migration rates
decreasing following formation of the first dune-track ridge.
Significant differences are observed between the dunes for
the interval between the formation of DT 5 and BS 6. Five of
the dunes appear to show a significant increase in migration
rate at this time, whereas one appears to show a dramatic
decrease. The explanation for these apparent differences
likely relates to the age of the back slope. Unlike dune-track
ridges, which record fluctuations in the groundwater table

During this early stage of the activity cycle, the morphology of the dune may more resemble a barchan than a parabolic dune, depending on the thickness of the surficial dry
active sand layer (David, 1993). Once the vegetation cover
has disappeared, the thick dry sand layer prevents more
humid sand at greater depth from drying out completely.
Owing to the absence of fine-grained particles, water can
only migrate upward through dry sand as vapour, and not by
capillary action (Prill, 1968). The absence of capillary movement in a bed of dry sand will allow it to remain dry after it has
been enclosed by humid sand (Kuhlman, 1957).

P.P. David et al.

A subsequent return to more humid conditions results in


moisture accumulation at the surface of the dune, and a
change from dry sand with transport-limited conditions to
wet sand with supply-limited conditions, coupled with partial
stabilization of the dune (Fig. 9, stage 4) and a consequent
abrupt reduction in migration rates (Fig. 8). Compared to dry
sand, moist sand requires higher wind speeds to initiate transport and has lower sediment transport rates due to the cohesive effects of the moisture (Namikas and Sherman, 1995).
The process of remoistening the sand in the whole dune is
a long-term, gradual one that occurs at the front of the dune,
where sand, moistened by occasional rain or the spring snow
melt, becomes buried by newly blown dry sand which then
protects it from drying out (Fig. 9, stage 5). In turn, the
surficial dry sand also becomes moistened and buried, resulting in the "backward" extension of humid sand in the whole

dune. At the same time, moist sand over the back slope of the
dune (Fig. 9, stage 4) is routinely dried out and transported
forward. As the dry sand supply becomes exhausted, the rate
of migration of the dune will drop dramatically, even without
any change in wind strength.
During the first return to less humid conditions, renewed
migration moves the head of the dune forward, leaving
behind a stabilized upwind portion of the dune base, forming
the first dune-track ridge (DT 2). Continued cycling of more
humid conditions accompanied by partial stabilization and
drought periods marked by renewed activity results in the formation of multiple track ridges (Fig. 9, stage 6). Following
formation of the final dune-track ridge, a slow process of
complete stabilization occurs, but may be periodically interrupted by more extensive sand-transport events in the
still-active portions of the dune.

neaativt? groundwater bi

Renewed activity

Stabilization of

. .

. -"

"

...

-.--

-.

/ I , '

\W,e.ak ItiI:e;gatige,,gyio.tha@M.atefilbI!@g;e$!

Figure 9. Schematic diagram showing stages of development of a back ridge and dune-track ridges during an
activity cycle of a parabolic dune. Stages 4 and 5 are repeated three times, with the newly wetted sand
expanding "backward" in the dune, before thefinal stage 6 is reached. The density and length of rays of the sun
symbol is a schematic representation of relative aridity, with phases 2 and 3 representing the most arid
intervals.

GSC Bulletin 534


Although dune activity may be initiated by human or animal activity in the absence of severe drought, we argue that
the record of the Seward sand hills shows a response to climatic fluctuations. Although there are many sites in the region
documenting activity by aboriginal hunters (David, 1993),
there are no indications of such activity in the immediate area
of the dunes studied here. In addition, where localized disturbance initiated activity, evidence would likely be ofrestricted
occurrence, small in extent and of limited intensity. For
example, present human disturbance tends to produce narrower point-source blowout depressions.
The sequence of ages obtained from the track ridges of
dune 2 provides supporting evidence for the dune activity
cycle presented above (Fig. 9, stage 6). At this dune, preservation of the ridges without significant subsequent reworking
occurred because of the comparatively low elevation of the
track ridges, resulting in nearness to the water table and lush
vegetation cover. In contrast, the track ridges of the larger
dune (dune 1) were probably more susceptible to deflation
and deposition in times of subsequent drought (Table 1).

During this activity cycle, dune migration was interrupted


by fluctuations in the water table four times, in each case producing a dune-track ridge. The final dune-track ridge predates the last major phase of advance when the dunes moved
to their present-day position. This event either reworked or
buried the back ridges of both dated dunes to a depth of more
than 50 cm. The present, partly stabilized dunes became
established by the end of the nineteenth century, with only
minor modifications of the ridges occurring since then. Final
stabilization of the heads of the dunes likely occurred at
slightly different times.

ACKNOWLEDGMENTS
The senior author (P.P.D.) is profoundly grateful to the
Geological Survey of Canada for having made this work possible within the Palliser Triangle Global Change Project. The
notion of dating dune-track ridges in the Seward sand hills
originated with the second author (S.A.W.). Dating of the
sands was done at the laboratory of the third author (D.J.H.),
who is grateful for financial support from the Natural Sciences
and Engineering Research Council of Canada (NSERC).

Finally, there is evidence that some of the dunes studied


underwent at least one earlier activity cycle prior to the one
documented in this study. One or two very low arcuate ridges,
resembling dune-track ridges, are discernible on aerial photographs upwind of, and parallel to, the modern back ridges.
The ridges are relatively wide and less than 20 cm high, and
hence inconspicuous in the field. It is likely that all the studied
dunes had similar past evolution, but that evidence of these
previous activity cycles has either been obliterated or overrun
by other dunes.

Figures were prepared by Michel Demidoff of the


UniversitC de MontrCal. The Dtpartement de gtologie of the
UniversitC de MontrCal defrayed the cost involved in the
preparation of the manuscript. Sonya Utting and Susan Ball
assisted with collection and preparation of samples. Sample
preparation and measurements for optical dating were carefully performed by G.O. Morariu.

SUMMARY

Comments by R.E. Vance, J.R. Prescott, O.B. Lian, and


R.G. Roberts, along with the formal reviews of V.T. Holliday
and an anonymous referee, helped improve the manuscript.

Field evidence from the Seward sand hills indicates that the
last dune-activity cycle started near the beginning of the nineteenth century and lasted until near the end of it. Sometime
before the cycle began, climate became more arid and produced a strongly negative groundwater budget. Consequently, moisture in the dune sands was depleted by the
stabilizing vegetation, although infrequent precipitation
allowed the vegetation to survive and resist wind action for
perhaps as long as 10-15 years. By about AD 1810, dune
migration began northeastward. Although it is unlikely that
all dunes began moving at precisely the same time, all were
probably active within a short time. The first phase of activity
was likely transport limited, with ample dry sand resulting in
rapid advance of the dunes. The abundance of dry sand may
have resulted in dunes assuming barchan morphology.
Increased precipitation during subsequent years began to
recharge regional groundwater. As the water table rose to
near, or above, ground level, the basal dune sands became
humid and vegetation covered the margins of the dunes. At
the same time, surficial dune sands became sufficiently
humid to provide increased resistance to wind action, producing supply-limited conditions similar to those that characterize the region today (Muhs and Wolfe, 1999). Following a
brief interval of limited activity, the dunes began migrating at
a reduced rate of approximately 2 m a 1 .

REFERENCES
Acton, D.F., Clayton, J.S., Ellis, J.G., Christiansen, E.A.,
and Kupsch, W.O.
1960: Physiographic divisions of Saskatchewan; Saskatchewan Research
Council, Geology Division, Map 1, scale 1:520 640.
Aitken, M.J. and Xie, J.
1992: Optical dating using infrared diodes: young samples; Quaternary
Science Reviews, v. 11, p. 147-152.
Arbogast, A.F.
1996: Stratigraphic evidence for late-Holocene aeolian sand rnobilizatio~l
and soil fonnation in south-cent~alKansas, U.S.A.;Journal of Arid
Environments, v. 34, p. 403414.
Berger, D.L.
1992: Ground-water recharge through active sand dunes in northwestern
Nevada; Water Resources Bulletin, v. 28, p. 959-965.
Borbwka, R.K.
1979: Accumulation and redeposition of eolian sands on the lee slopes of
dunes and their influence on formation of sedimentary structures;
Quaestiones Geographicae, v. 5, p. 5-52.
Clarke, M.L.
1994: Infra-red stimulated luminescence ages from aeolian sand and alluvial fan deposits from the eastern Mojave Desert, California;
Quaternary Science Reviews, v. 13, p. 533-538.
David, P.P.
1964: Surficial geology and groundwater resources of the Prelate area
(72 K), Saskatchewan; Ph.D, thesis, McGill University, MontrCal,
QuBbec, 329 p.

P.P. David et al.

David, P.P. (cont.)


1968: Geomorphology, stratigraphy, chronology and migration of sand
dunes in Manitoba and Saskatchewan; in Report of Activities; Part A,
Geology Survey of Canada, Paper 68-lA, Part A, p. 155-157.
1969: Sand dunes studies in southwestern Saskatchewan (72K); in Report
of Activities, Part A; Geological Survey of Canada, Paper 69-1, Part
A, p. 61-62.
1971: TheBrookdaleroad section and its significancein the chronological
studies of dune activities in the Brandon sand hills of Manitoba;
Geological Association of Canada, Special Paper 9, p. 293-299.
1972: Great Sand Hills, Saskatchewan,Day 7 -September 6; in Quaternary Geology and Geo~norphology between Winnipeg and the
Rocky Mountains, (ed.) N.W. Rutter and E.A. Christiansen; 24"'
International Geological Congress, Montreal, Quebec, Guidebook
for Excursion C-22, p. 36-50.
1981: Stabilized dune ridges in northern Saskatchewan;Canadian Journal
of Earth Sciences, v. 18, p. 286-310.
1988: The coeval eolian environment of the Champlain Sea Episode; in
The Late Quaternary Development of the Champlain Sea Basin,
(ed.) N.R. Gadd; Geological Association of Canada, Special Paper
35, p. 291-305.
1993: Great Sand Hills of Saskatchewan:an overview; in Quaternary and
Tertiary Landscapes of Southwestern Saskatchewan and Adjacent
Areas, (ed.) D.J. Sauchyn; Canadian Plains Research Center,
University of Regina, ~ e s n aSaskatchewan,
,
Field Trip Guidebook
for the INQUA Commission on Formation and Propertiesof Glacial
Deposits Field Conference, 1993, p. 59-81.
Dobrovolny, E. and Townsend, R.C.
1947: An eolian dam near Tucumcari, New Mexico; Economic Geology
and the Bulletin of the Society of Economic Geologists, v. 42,
p. 492-497.
Duller, G.A.T.
1994: Luminescence dating using feldspars: a test case from southern
North Island, New Zealand; Quaternary Science Reviews, v. 13,
p. 423-427.
Environment Canada
1993: Canadian climate normals, 1961-1990: Volume 2, Prairie
Provinces; Minister of Supply and Services, 266 p.
Finkel, H.H.
1959: The barchans of southern Peru; Journal of Geology, v. 67,
p. 614-647.
Forman, S.L., Oglesby, R., Markgraf, V., and Stafford, T.
1995: Paleoclimatic significance of late Quaternary eolian deposition on
the Piedmont and High Plains, central United States; Global and
Planetary Change, v. 11, p. 35-55.
Fryberger, S.G., Al-Sari, A.M., Clisham, T.J., Rizvi, S.A.R.,
and AI-Hinai, K.G.
1984: Wind sedimentation in the Jafurah sand sea, Saudi Arabia;
Sedimentology, v. 31, p. 413-431.
Fuller, I.C., Wintle, A.G., and Duller, G.A.T.
1994: Test of the partial bleach methodology as applied to the infra-red
stimulated luminescence of an alluvial sediment from the Danube;
Quaterna~yScience Reviews, v. 13, p. 539-543.
Gupta, J.P.
1979: Some observations on the periodic variations of moisture in
stablised and unstabilised sand dunes of the Indian Desert; Journal
of Hydrology, v. 41, p. 153-156.
Halsey, L.A., Catto, N.R., and Rutter, N.W.
1990: Sedimentologyand developmentof parabolic dunes, Grande Prairie
dune field, Alberta; Canadian Journal of Earth Sciences, v. 27,
p. 1762-1772.
Hastenrath, S.L.
1967: Thc barchans of the Arequipa region, southern Peru; Zeitschrift fur
Geomo~phologie,v. 11, p. 300-33 1.
Haynes, C.V., Jr.
1989: Bagnold's barchan: a 57-yr record of dune movement in the eastern
Sahara and implications for dune origin and paleocljmate since
Neolithic times; Quaternary Research, v. 32, p. 153-167.
Holliday, V.T.
1995: Late Quaternary stratigraphy of the Southern High Plains; in
Ancient Peoples and Landscapes, (ed.) E. Johnson; Museum of
Texas Tech University, Lubbock, Texas, p. 289-313.
Hunter, R.E.
1985: A kinematic model for the structure of lee-side deposits;
Sedimentology, v. 32, p. 409-422.

Huntley, D.J. and Clague, J.J.


1996: Optical dating of tsunami-laid sands; Quaternaly Research, v. 46,
p. 127-140.
Huntley, D.J. and Lian, O.B.
1999: Using optical dating to determinewhen a sediment was last exposed
to sunlight; in Holocene Climate and Environmental Change in the
Palliser Triangle: A Geoscientific Context for Evaluating the
Impacts of Climate Changeon the Southern Canadian Prairies, (ed.)
D.S. Lemmen and R.E. Vance; Geological Survey of Canada,
Bulletin 534.
Hiitt, G., Jaek, I., and Tchonka, J.
1988: Optical dating: K-feldspars optical response stimulation spectra;
Quaternary Science Reviews, v. 7, p. 381-385
Jennings, J.N.
1965: Further discussion of factors affecting coastal dune formation in the
tropics; Australian Journal of Science, v. 28, no. 4, p. 166-167.
Keen, K.L.
1987: Groundwater recharge from a summer precipitation event in a stabilized sand dune environment, western Nebraska Sand Hills;
Geological Society of America, Abstracts with Programs, v. 19,
p. 207.
Keen, K.L. and Rosenberry, D.O.
1988: Rapid focused-recharge at a stabilized parabolic dune in the
Nebraska Sand Hills (Crescent Lake National Wildlife Refuge);
EOS, v. 69, no. 16, p. 354.
Kerr, R.C. and Nigra, J.O.
1952: Eolian sand control; American Association of Petroleum
Geologists Bulletin, v. 36, p. 1541-1573.
Kuhlman, H.
1957: Sandflugt og klitdannelse; Geografisk Tiddskrift, v. 56, p. 1-19.
Lanner, D.
1992: Discovery of dune dams reveals formation of sand hills lakes;
University of Lincoln, Lincoln, Nebraska, Conservationand Survey
Division, Research Notes 6, p. 9-1 1.
Lemmen, D.S., Vance, R.E., Campbell, I.A., David, P.P.,
Pennock, D.J., Sauchyn, D.J., and Wolfe, S.A.
1998: Geomorphic systems of the Palliser Triangle, southern Canadian
Prairies: description and response to changing climate; Geological
Survey of Canada, Bulletin 521, p. 25-39.
Lian, O.B.
1997: Quaternary geology of the Fraser Valley area, Big Bar Creek to
Pavilion, south-central British Columbia; Ph.D. thesis, University
of Western Ontario, London, Ontario, 489 p.
Lian, O.B., Hu, J., Huntley, D.J., and Hicock, S.R.
1995: Optical dating studies of Quaternary organic-rich sediments from
southwestern British Columbia and northwestern Washington
State; Canadian Journal of Earth Sciences, v. 32, p. 1194-1207.
Loewy, H.
1953: Senkwasser in Sandwiisten; Neues Jahrbuch fiir Geologie und
Palaeontologie, Monatschefte 6, p. 241-243.
Loope, D.B., Swinehart, J.B., and Mason, J.P.
1995: Dune-dammed paleovalleys of the Nebraska Sand Hills: intrinsic
versus climatic controls on the accumulationof lake andmarsh sediments; Geological Society of AmericaBulletin, v. 107,p. 396-406.
Madole, R.F.
1994: Stratigraphic evidence of desertification in the west-central Great
Plains within the past 1000 yr; Geology, v. 22, p. 483486.
McCord, J.T. and Stephens, D.B.
1987: Effect of ground-water recharge in configuration of the water table
beneath sand dunes and on seepage in lakes in the sandhills of
Nebraska, U.S.A. - comment; Journal of Hydrologv,
-. v. 95,
p. 365-367.
McKee., E.D.., Dou~lass.-1.R.. and Rittenhouse. S.
1971: ~eformationoflee-side laminae in eolian dunes; Geological
Society of America Bulletin, v. 82, p. 359-378.
Mejdahl, V.
1979: Thermoluminescence dating: beta-dose attenuation in quartz
grains; Archaeometry, v. 21, p. 61-72.
MolnAr, B.
1961: A Duna-Tisza kozi eolikus rktegek felszini 6s felszln alatti
kiterjedese; Foldtani K6zlony, v. 91, p. 300-315.
Muhs, D.R. and Holliday, V.T.
1995: Evidence of active dune sand on the Great Plains in the 19" century
from accounts of early explorers; Quaternaly Research, v. 43,
p. 198-208.

GSC Bulletin 534


Muhs, D.R. and Wolfe, S.A.
1999: Sand dunes of the northern Great Plains of Canada and the United
States; in Holocene Climate and Environmental Change in the
Palliser Triangle: A Geoscientific Context for Evaluating the
Impacts of Climate Change on the Southern Canadian Prairies, (ed.)
D.S. Lemmen and R.E. Vance; Geological Survey of Canada,
Bulletin 534.
Muhs, D.R., Stafford, T.W., Jr., Been, J., Mahan, S., Burdett, J.,
Skipp, G., and Rowland, Z.M.
1997a: Holocene eolian activity in the Minot dune field, North Dakota;
Canadian Journal of Earth Sciences, v. 34, p. 1442-1459.
Muhs, D.R., Stafford, T.W., Jr., Cowherd, S.D., Mahan, S.A.,
Kihl, R., Maat, P.B., Bush, C.A., and Nehring J.
1996: Origin of the late Quaternary dune fields of northeastern Colorado;
Geomorphology,v. 17, p. 129-149.
Muhs, D.R., Stafford, T.W., Jr., Swinehart, J.B., Cowherd, S.D.,
Mahan, S.A, Bush, C.A., Madole, R.F., and Maat, P.B.
1997b: Late Holocene eolian activity in the ~nineralogically mature
Nebraska Sand Hills; Quaternary Research, v. 48, p. 162-176.
Nambi, K.S.V. and Aitken, M.J.
1986: Annual dose conversion factors for TL and ESR dating;
Archaeometry, v. 28, p. 202-205.
Namikas, S.L. and Sherman, D.J.
1995: A review of the effects of surface moisture content on aeolian sand
transport; in Desert Aeolian Processes, (ed.) V.P. Tchakerian;
Chapman and Hall, New York, p. 269-293.
Ollerhead, J., Huntley, D.J., and Berger, G.W.
1994: Luminescence dating of sediments from Buctouche Spit, New
Brunswick; Canadian Journal of Earth Sciences, v. 31 ,p. 523-53 1.
Phillips, J.D., Perry, D., Garbec, A.R., Carey, K., Stein, D.,
Morde, M.B., and Sheehy, J.A.
1996: Deterministic uncertainty and complex pedogenesis in some
Pleistocene dune soils; Geoderma, v. 72, p. 147-164.
Prescott, J.R. and Hutton, J.T.
1988: Cosmic ray and gamma ray dosimetry for TL and ESR; Nuclear
Tracks and Radiation Measurements, v. 14, p. 223-227.
Prescott, J.R. and Robertson, G.B.
1997: Sediment dating by luminescence: a review; Radiation
Measurements, v. 27, p. 893-922.

Prill, R.C.
1968: Movement of moisture in the unsaturated zone in adunearea, southwestern Kansas; United States Geological Survey, Professional
Paper 600, p. Dl-D9.
Pye, I<.
1980: Beach salcrete and eolian sand transport: evidence from North
Queensland;Journal of SedimentaryPetrology, v. 50, p. 257-26 1.
Spooner, N.A., Aitken, M.J., Smith, B.W., Franks, M.,
and McElroy, C.
1990: Archaeologicaldating by infrared-stimulatedluminescenceusing a
diode array; Radiation Protection Dosimetry, v. 34, p. 83-86.
St-Onge, D.A.
1966: Geomorphology of the Lancer area, Saskatchewan; Revue de
GBographie de Montreal, v. 20, p. 2 7 4 5 .
Wheaton, E.E. and Chakravarti, A.K.
1987: Some temporal, spatial and climatological aspects of dust storms in
Saskatchewan; Climatological Bulletin, v. 21, p. 5-16.
Winter, T.C.
1986: Effect of ground-water recharge in configuration of the water table
beneath sand dunes and on seepage in lakes in the sandhills of
Nebraska, U.S.A; Journal of Hydrology, v. 86, p. 221-237.
Wolfe, S.A.
1997: Impact of increased aridity on sand dune activity in the Canadian
Prairies; Journal of Arid Environments, v. 36, p. 421432.
Wolfe, S.A. and David, P.P.
1997: Canadian landforms example: parabolic dunes; Canadian
Geographer, v. 41, p. 207-213.
Wolfe, S.A and Lemmen, D.S.
1999: Monitoring of sand dune activity in the Great Sand Hills region,
southwestern Saskatchewan; iiz Holocene Climate and Environmental Change in the Palliser Triangle: A GeoscientificContext for
Evaluating the Impacts of Climate Change on the Southern
Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological
Survey of Canada, Bulletin 534.
Wolfe, S.A., Huntley, D.J., and Ollerhead, J.
1995: Recent and late Holocene sand dune activity in southwestern
Saskatchewan; in Current Research 1995-B; Geological Survey of
Canada, p. 131-140.

Geomorphology of the western Cypress Hills:


climate, process, stratigraphy, and theory
David J. sauchynl
Sauchyn, D.J., 1999: Geomorphology of the western Cypress Hills: climate, process,
stratigraphy, and theory; in Holocene Climate and Environmental Change in the Palliser
Triangle: A Geoscientific Contextfor Evaluating the Impacts of Climate Change on the Southern
Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological Survey of Canada,
Bulletin 534, p. 239-247.

Abstract: This paper relates the stratigraphic record from the western Cypress Hills to observations of
process and the theoretical framework of process geomorphology. Adjustments of this landscape to changing boundary conditions (climate and resistance) are triggered by hydroclimatic events but conditioned by
the history and inherent instability of geomorphic systems. The notion that geology (resistance) is an independent variable is contrary to the geomorphic history. Quasi-continuous processes predispose the landscape to higher magnitude, lower frequency events. Conversely, infrequent deep-seated slope failures
impact stream geometry and sediment load. A progressive decline in lake sedimentation conforms to an
equilibrium model of watershed geomorphology, involving increasing morphological and filter resistance.
This complex and episodic response to climatic change and variability should result in systematic differences among stratigraphic records. The sediments shed from increasingly larger areas should record climatic forcing of geomorphic events of increasing magnitude and decreasing frequency.

Resum6 : Cet article associe les donnCes stratigraphiques de la partie ouest des collines Cypress & des
observations sur les processus et au cadre thkorique de la gkomorphologie des processus. Les ajustements
de ce paysage aux changements de conditions aux limites (climat et rCsistance) sont dCclenchCs par des
kvCnements hydroclimatiques, mais conditionnks par I'historique et par l'instabilitk intrindque des
systkmes gComorphologiques. L'idCe que la gCologie (rCsistance) est une variable indCpendante est contraire & l'histoire gComorphologique. Des processus quasi-continus pr6disposent le paysage B des
CvCnements de forte magnitude et de faible frkquence. Inversement, de rares ruptures de pente d'origine
profonde se repercutent sur la gComCtrie des cours d'eau et leur charge sedimentaire. Une diminution progressive de la skdimentation lacustre est conforme A un modkle d'kquilibre de la gtomorpologie des bassins
versants mettant en jeu une resistance accrue d'origine morphologique et par filtrage. Cette rkponse
complexe et kpisodique aux changements et la variabilitk climatiques devrait se traduire par des
diffkrences systkmatiques parmi les donnCes stratigraphiques. Dans la mesure oii les sCdiments sont issus
de rkgions de plus en plus Ctendues, ils devraient reflCter un f o r ~ a g eclimatique d'kvtnements
gComo~phologiquesde magnitude croissante et de frkquence dkcroissante.

Department of Geography, University of Regina, Regina, Saskatchewan S4S 0A2

GSC Bulletin 534

INTRODUCTION
With the priority given to global change research, climatic
change and variability have become a focus for regional studies of geomorphic processes (e.g. Thomas and Allison, 1993;
Jones, 1993). Environmental concerns require the study of
larger areas and time spans to establish environmental variability and the impact of human activities (Vitek and
Giardino, 1993),challenging geomorphologists to apply their
understanding of processes to landscapes and the sedimentary record. Research associated with the Palliser Triangle
Global Change Project has attempted to evaluate linkages
between Holocene climate and regional geomorphic processes (Lemmen et al., 1993). The southern Interior Plains are
well suited to such studies, given their tectonic stability and
the sensitivity of subhumid landscapes to changes in the climatically driven surface water balance (Bull, 1991;
Campbell, 1997).The project objectives, vast study area, and
a previous lack of a paleogeomorphic data favour the study of
stratigraphic records, integrating geomorphic responses to
climate over time (e.g. Wolfe et al., 1995), as opposed to the
monitoring or modelling of process.
The inference of climatic forcing of geomorphic process
from sediments and sedimentary structures requires a convergence of methodologies, which have been historically segregated among the earth sciences despite a common goal of
reconstructing process and landscape. "The very nature of
landscapes requires geomorphology to assume the dual
nature of being both a historical and physical science."
(Ritter, 1988,p. 160).Process geomorphology is based on the
explanation of landform and sediment yield in terms of process mechanics, the concept of the geomorphic system, and
the time independence of geomorphic activity, as a function
of spatial variation in controls and resistance, and the mutual
adjustment of process and form. In contrast, studies of landscape evolution are naturally historical, with the emphasis on
stratigraphy, chronology (time dependence), age correlation
of lithofacies, and reconstruction of sedimentary
environments.
This paper attempts to apply both perspectives to an interpretation of proxy and process data from the western Cypress
Hills of southwestern Saskatchewan and southeastern
Alberta (Fig. 1). Much of these data are reported elsewhere
(e.g. Goulden and Sauchyn, 1986; Sauchyn, 1990), including
this volume (Sauchyn and Nelson, 1999; Spence and
Sauchyn, 1999). Here they are re-examined in a theoretical
context of geomorphic response to climatic change and variability. A previous overview of the geomorphology of the
Cypress Hills (Sauchyn, 1993a) preceded research in conjunction with the Palliser Project and made no attempt to
relate field observation to geomorphic theory.

BACKGROUND
The primary evidence for the reconstruction of geomorphic
history is fragmentary, discontinuous, and possibly unrepresentative deposits (Clayton, 1983). Whereas paleosols are the
geological record of 'stable' landscapes, lithostratigraphy

represents the periodic response of geomorphic systems to


climatic change, extreme hydroclimatic events, intrinsic
thresholds, andlor lowered resistance. The timing and duration of this response can be out of phase with the climate
associated with the process domain, while the disequilibrium
propagates through the geomorphic system. For example,
glacial deposits, landforms, and tectonic structures have a
prolonged influence on the apparent geomorphic response to
postglacial climates (Church and Ryder, 1972; Matheson and
Thornson, 1973; Sauer, 1978; Church and Slaymaker, 1989;
Vreeken, 1999).
Process geomorphology and sedimentary geology are
complementary approaches to the study of landscapes.
Subaerial process is ubiquitous; terrestrial sedimentation is
spatially discontinuous and time transgressive. "The
sedimentologist and stratigrapher have great temporal freedom.. .the geomorphologist, on the other hand, has spatial
freedom." (Schurnrn, 1981, p. 19). Stratigraphic data, the
cumulative record of geomorphic events, enable the testing of
hypotheses derived from a theoretical understanding of
geomorphic systems. Attention to the stratigraphic record
and geological history prevents process models from explaining inherited landforms in terms of contemporary process.

res

Figure 1. A) Location of the western Cypress Hills. Palliser


Triangle delimited by dashed line, light shading denotes
Brown Chernozemic Soil Zone (see Lemmen et al., 1998),and
the three blocks of the Cypress Hills are shaded dark.
B) Study sites mentioned in text. Upper plateau sugace is
shaded.

D.J. Sauchyn

Conversely, the study of modern environments and processes


enables facies modelling (Walker, 1984) and "saves" sedimentary geology from being simply ". . .the barren cataloguing of rock and fossil sequences that sometimes passes for
stratigraphy" (Middleton, 1973, p. 983).
Understanding of geomorphic processes is derived both
from field experiments, which are incapable of explaining
landscape evolution (Campbell, 1982; Eybergen and Imeson,
1989), and from models, which have yet to emulate the
observed complexity and range of scales of geomorphic process (Baker and Twidale, 1991). Current geomorphic theory
developed as a conceptual framework for process geomorphology, particularly the study of fluvial processes (Ritter,
1988). An obsession with theory, however, to satisfy scientific rigour can detach process geomorphology from real
landscapes. "There may be a point at which we are no longer
able to understand our model any better than the landscape
which it simulates." (Kirkby, 1990,p. 79). A sirnilarpreoccupation with methodology exposes stratigraphic research to a
tendency to uncritically perpetuate geochronologies that may
be more believed than real (Bowen, 1978; Davies, 1989;
Ager, 1993).
Notwithstanding facies modelling and stratigraphic principles, sediment studies tend to be inductive. "Quaternary science tends to emphasize the role of the 'historical-inferential'
approach over that of the theoretical." (Thorn, 1988, p. 220).
Observation and theory often are presented and debated as
opposing philosophies (Baker and Twidale, 1991; Rhoads
and Thorn, 1993).In practice, however, observation is guided
by preconception, if not theory, and theory has no purpose or
basis without observation. Only those questions with a theoretical basis can lead to universal explanation (Rhoads and
Thorn, 1993). Thus, facies and process modelling (Kirkby,
1990; Walker and James, 1992) are required to link
short-term process observation to the stratigraphic record,
and to interpret geomorphic response to climate from differentially preserved sediments.

CLIMATE, STRATIGRAPHY, AND PROCESS


The Palliser Triangle includes a variety of landscapes (e.g.
dunes fields, badlands, ice-thrust hills) with distinctive
geomorphic histories and responses to climatic variability
and change (Sauchyn, 1993b; Leinmen et al., 1998). In the
Cypress Hills (Fig. I), "the hub of the Palliser Triangle"
(Vance and Last, 1994, p. 51), coniferous forest and several
hundred metres of local relief contrast sharply with the surrounding mixed-grass prairie and glaciated plains. The same
geological units, however, underlie both the Tertiary sediments which cap the Cypress Hills and the Pleistocene
deposits on the surrounding plains. Thus there is geological
continuity, unlike the nearby uplands of tectonic origin in
north-central Montana.
The physiography and geomorphic history of the western
Cypress Hills provide an opportunity for the study of
geomorphic responses to climate at various scales. Unlike
most of Canada, the palimpsest which is the western Cypress

Hills was not overwhelmed by Laurentide glaciation, but


rather formed a 300 km2 nunatak with some 90 m of relief
(Stalker, 1965). This paper focuses on the landscape beyond
the limit of Wisconsinan glaciation (Klassen, 1994), where
there is stratigraphic and morphological evidence of fluvial,
glaciofluvial, mass wasting, and eolian processes operating
since the late Tertiary (Sauchyn, 1993a), and significant contemporary geomorphic activity (Sauchyn and Nelson, 1999;
Spence and Sauchyn, 1999). A Holocene climatic history has
been established based on a well controlled sedimentary
record from Harris Lake (Fig. 1;Sauchyn, 1990; Sauchyn and
Sauchyn, 1991; Last and Sauchyn, 1993; Wilson et al., 1997;
Porter et al., 1999) and from the dendroclimatology of the
spruce-pine forest (Sauchyn and Beaudoin, 1998).
The Cypress Plain, the highest and oldest surface in the
Interior Plains, is developed in Eocene to mid-Miocene fluvial gravel deposits. Leckie and Cheel(1989) concluded that
these sediments are remnants of a paleo-braidplain deposited
during a semiarid climatic regime in response to the Eocene
emergence and denudation of the ancestral Rocky Mountains, and subsequent intrusion of the Sweet Grass Hills. With
a change to a wetter climate, the ancestral Missouri and
Saskatchewan river systems degraded the interior sedimentary basin, leaving only the continental divide (i.e. the
Cypress Hills). This scenario inspired Crickmay's (1932,
1975) concept of unequal activity, a much neglected principle
"with major implications for landscape evolution." (Twidale,
1993,p. 357). He proposed geological control for the survival
of the Cypress Hills. The permeable caprock resists erosion
by readily conducting water or evolving to a resistant rimrock
of cemented sand and gravel (Kupsch and Vonhoff, 1967).
Tertiary landforms have resisted change and thus dominate
the gross morphology of the Cypress Hills. The age of tephra
embedded in loess indicates that the erosional surfaces had
formed by 8.3 Ma (Vreeken and Westgate, 1992). Jungerius
(1967) attributed the deposits and low relief on the pediments, extending from the Cypress Plains to the present valley network, to intervals of cyroplanation and pediplanation.
Other relict landforms are the late Pleistocene meltwater valleys, which circumscribe and bisect the west block, and are a
locus of late Quaternary geomorphic activity (cf.
Christiansen and Sauer, 1988).
The most apparent geomorphic responses to Holocene
climate coincide with the onset and waning of maximum
Holocene temperature and aridity. Prior to the Hypsithermal
(ca. 7500-5000 BP; Sauchyn and Sauchyn (1991), Porter
et al. (1999), a relatively stable landscape is inferred from
strongly developed paleosols in southern Alberta (Waters and
Rutter, 1984; Valentine et al., 1987), and low sedimentation
rates and a relative abundance of arboreal (aspen) pollen in
Harris Lake (Sauchyn and Sauchyn, 1991). Biostratigraphic
and lithostratigraphic records that postdate deposition of
Mazama ash (6800 BP; Bacon (1983)) signal a warmer, drier
climate, reduced vegetation cover, and disequilibrium in the
fluvial and eolian systems. During the interval from
6100-5300 BP, sedimentation significantly increased in
Harris Lake and organic matter declined to less than 7%, indicating higher yields of mineral sediment andlor lower plant
productivity (Sauchyn, 1990). A unit of mid-Holocene

GSC Bulletin 534

alluvium(Westgate, 1972)andpost-Mazamaloess (Vreeken,


1986) is further evidence of accelerated erosion and
sedimentation.
The cooler and wetter post-Hypsithermal was a period of
landscape stability and soil formation in the southern Interior
Plains (e.g. Valentine et al., 1987). Previous authors concluded that floodplains and alluvial fans in the Cypress Hills
were likewise relatively stable (Jungerius 1969; Jungerius
and Mucher, 1969; Westgate, 1972). More recent research,
however, points to significant slope instability after about
4500 BP (Goulden and Sauchyn, 1986; Sauchyn, 1990; Last
and Sauchyn, 1993; Sauchyn and Lernmen, 1996).The wetter
climate prompted a readjustment of valley sides, as water
tables rose in response to increased groundwater recharge and
decreased evapotranspiration. The absolute and relative ages
of 21 landslides in this area are all post-Hypsithermal
(Sauchyn and Lernmen, 1996). In Harris Lake, 5120-3450
BP was the period of maximum sedimentation. A dramatic
shift from mainly detrital to mainly endogenic sediments,
after ca. 4000 BP, can be linked to a landslide that closed the
lake basin (Last and Sauchyn, 1993; Porter et al., 1999). Further landsliding in the watershed caused striking fluctuations
in the ratio of siliciclastic to endogenic carbonate sediment.
Although the lacustrine and landslide stratigraphy mostly
record the reworking of older landslides, these data, plus evidence for soil development on alluvial deposits, reflect forestation of the cypress Hills and an associated shift in dominant
geomorphic activity from erosion to landsliding.
Geomorphic activity associated withrain and snowmeltwater
is largely controlled by the permeability of the caprock and
variable hydraulic conductivity of the underlying strata.
Groundwater exfiltrating from stream banks, concave slopes,
and valley heads accounts for much of the current geomorphic
activity (Spence and Sauchyn, 1999).The two largest historical geomorphic events occurred when more than 1.5 m of
snow fell in late April and early May 1967, and temperatures
rose dramatically on May 14 (Janz and Treffry, 1968). Snow
meltwater triggered both Police Point landslide (Sauchyn and
Nelson, 1999) and flood discharge on nearby Graburn Creek
with areturn period of greater than 50 years (MacPherson and
Rannie, 1969). Casual and systematic field observations
(Sauchyn and Lemmen, 1996; Sauchyn and Nelson, 1999)
suggest that, during drought years, large available soil-water
storage and adequate vegetation cover prevent surface erosion, while significant erosion and mass wasting occur with
above-average annual precipitation. Throughout the western
Cypress Hills, there is dendrogeomorphic evidence of flooding (Boon, 1993) and slope movement (Josephson, 1988) of
sufficient magnitude to disturb tree growth. Spruce seedlings
that colonized Police Point landslide during the drier mid- to
late 1980s have been undermined and uprooted as a result of
increased instability during recent wetter years.

THEORY
The complex and nonlinear (threshold) behaviour of
geomorphic systems is described by a body of concepts that
serve as a theory of geomorphic response to climatic change

and variability (Schurnm, 1975, 1977, 1981; Thornes and


Brunsden, 1977; Brunsden and Thornes, 1979; Brunsden,
1990, 1993; Bull, 1991). A geomorphic response consists of
reaction time, the interval between a disturbance or change in
controls and observable morphological change, plus relaxation time, the ensuing period of adjustment up to a new equilibrium or another threshold (Fig. 2; Bull, 1991). Whereas
climatic change tends to produce a regional response and,
ultimately, shifts in process domains, hydroclimatic events
(climatic variability) and internal perturbations (intrinsic
thresholds) generally are more difficult to recognize or verify
stratigraphically. Rather than a response to climatic forcing,
sediments can represent the self-regulation of a geomorphic
system (dynamic equilibrium), with exponentially decreasing activity after a period of more rapid adjustment, typically
involving negative feedback (Welch, 1970; Thornes and
Blunsden, 1977). A key variable in the reconstruction of
geomorphic history is the length of relaxation time relative to
the periodicity of disturbances or changes in controls which
exceed thresholds (Wolman and Gerson, 1978). Highmagnitude, low-frequency events tend to be overrepresented
both stratigraphically and morphologically, because the outcomes are persistent, unlike the typically more subtle
imprints of quasi-continuous erosion and sediment transport,
even though these processes may have a greater product of
magnitude and frequency.
Hydroclimatic events are the principal agents of
geomorphic activity in the dry, tectonically stable Palliser
Triangle. Their effectiveness depends largely on resistance as
a function of surficial geology, geomorphic history, and the
climatic control of vegetation and subsurface water. Differential geomorphic activity commonly reflects the distribution
Relaxation!

!
. .

Persistence

rn

.. .

Figure 2. Definition of response (R,) and persistence (P,)


times, in terms thresholds (T)and equilibria ( E ) in the history
of a geomorphic system (e.g. changes in streambed
elevation). Response time is the sum of reaction time (R,),
which is the interval between the perturbation (P)of a system
and observable change (threshold, T), and relaxation time
(R,), which is the subsequent interval until the next threshold
or equilibrium is established. Persistence (P,) is the duration
of the new equilibrium (modified from Fig. 1.3ofBull, 1991).

D.J. Sauchyn

of resistance and available sediment, or the unique histories


of geomorphic systems, rather than spatial variation in the
efficacy of climate as a geomorphic agent. Where resistance
is lacking, deposits are reworked and landforms modified,
and the stratigraphic record can be biased towards more
recent events (Clayton, 1983). Unusually thick sequences of
sediment can represent a local lack of resistance or an atypical
depositional environment (e.g. buried valleys). These sections attract the Quaternary geologist, in the same way that
process geomorphologists monitor the most active, but often
anomalous, landforms.
Most landscapes also include elements that resist change,
by virtue of geological properties (strength resistance) and/or
previous evolution to a stable form (filter and system-state
resistance; Brunsden (1993)). An example from the Palliser
Triangle is the mid-Holocene accumulation of loess, which
initiated a period of reduced erosion because infiltration rates
were higher in the loess than in the underlying Cretaceous
shales (Bryan et al., 1987). Structural resistance to landscape
change includes disorder in hydrological and geomorphic
systems. In the recently glaciated landscapes of the Palliser
Triangle, significant geomorphic activity in dune fields,
cropland, and large meltwater valleys tends to be isolated by a
lack of permanent streamflow and a poorly integrated drainage network (Lemmen et al., 1998). AS a result,~ratesof erosion on cropland are 2-3 times higher than sediment yields in
small watersheds (Carson and Associates, 1990).
Finally, a major contribution of process geomorphology
towards the understanding of landscape evolution is the recognition that the apparent behaviour of geomorphic systems
is relative to scales of space and time. Relationships among
controls and responses, and types of equilibrium, differ
among cyclic (geological), graded, and steady (engineering)
time (Schurnm and Lichty, 1965; Thornes and Brunsden,
1977).Processes operating at one spatial scale can predispose
a landscape to processes of different magnitudes and frequencies operating at another scale (deBoer, 1992). In the Alberta
badlands, for example, geomorphic thresholds are scale
dependent, such that the controls on sediment yield from basins are not evident or relevant at the scale of slopes (deBoer
and Campbell, 1989). This implies that the inference of process from sediments also requires recognition of type and
degree of coupling (structural resistance) between parts of a
geomorphic system: slopes, channels, and sediment sinks,
and, in general, discontinuous sedimentary deposits and the
erosional history of source areas (Schurnm, 1981). Disturbance of geomorphic systems tends to result in complex
sequences of erosional and depositional events, such that the
response of adjacent channels, slopes, or even cross-sections
can be expressed at different rates, and even at different
times. This complex response (Schurnm and Parker, 1973;
Womack and Schumm, 1977) is an aspect of filter resistance
(Brunsden, 1993) that causes changes in rates of erosion and
sedimentation independent of climatic variation.
Scientific philosophy dictates that "...it is impossible to
do anything more than trivial work without some theoretical
context." (Kirkby, 1990,p. 64). Abstract thoughts about landscapes can range from unconscious preconceptions to models
deduced from physical principles. Facies models are a

theoretical context for the universal interpretation of process


and environment from the stratigraphy and structure of sedimentary deposits (Walker, 1984), such as the Cypress Hills
Formation (Leckie and Cheel, 1989). Here, geomorphic systems theory is applied to an interpretation of stratigraphic and
process observations from the western Cypress
Hills.
-If recognition and understanding of landscape change are
scale dependent, then scales of space and time must be established for the western Cypress Hills. The complete
geomorphic history, late Holocene climatic and landscape
change, and individual hydroclimatic events occupy, respectively, cyclic, graded, and steady time. Space can be classified
using the obvious hierarchy of watersheds, valleys, and
slopes and channels. Alternatively, it can be classified
according to specific thresholds in morphology and process:
the limit of late Pleistocene glaciation (Klassen, 1994; Kulig,
1996); drainage divides, including the continental divide; the
edge of the Cypress Plain, separating the oldest surface in the
Interior Plains from the immediately adjacent landslide
scarps which are among the youngest and most active surfaces; the floodplain buffer between stream channels and the
valley sides almost universally dominated by landsliding; the
junctions between meltwater valleys and the tributary valleys
of fluvial origin; and the valley heads, where geomorphic
activity is dominated by seepage erosion and sapping during
the exfiltration of groundwater. These spatial thresholds, or
discontinuities between geomorphic systems, heavily influence the transmission of energy, material, and landscape
change, and therefore structural and filter resistance to disturbance and environmental change.
The Tertiary morphology and Quaternary stratigraphy of
the western Cypress Hills suggest a correlation of process and
paleoclimate, as process domains shifted among Pleistocene
periglaciation, subhumid fluvial and eolian processes, and
more humid mass wasting and fluvial erosion during
interglacials (including the Holocene). This apparent correlation over cyclic time reflects not only the low resolution of
long proxy records, but also landscape instability in response
to arid and periglacial paleoclimate, extremes of moisture and
temperature which are absent from the Holocene, except perhaps for the period of maximum warmth and aridity. The climatic changes away from aridity and periglaciation left
profound erosional imprints: the original downcutting of the
Cypress Hills and erosion of the meltwater valleys.
The tempo and more detailed record of climatic change
over graded time favour the recognition of climatic forcing of
geomorphic events. Given the relatively cool and moist late
Holocene climate, the stratigraphic record of climatic variability is mostly flood and landslide deposits, rather than the
wind-blown deposits of drier climates (e.g. Vreeken, 1993).
Low-magnitude, high-frequency processes, weathering and
the transport of dissolved solids in particular, are linked
directly to climate. In the Cypress Hills, however, rates of
weathering have little bearing on the timing and intensity of
geomorphic responses to climate. Poorly consolidated bedrock and hundreds of metres of local relief predispose this
landscape to landsliding and fluvial erosion. Sediment production is erosion and transport limited by the magnitude, frequency, and effectiveness of hydroclimatic events. Both

GSC Bulletin 534

extreme weather and resistance to it are linked to climate.


Strength resistance depends on pore-water pressure and vegetation cover (i.e. antecedent conditions).
Morphological and structural resistance develop as
geomorphic systems adjust to changing climate and extreme
events. Low-magnitude, high-frequency processes locally
modify landforms and sediments, thereby increasing their
resistance to further erosion of similar intensity. Exposure of
the sand and gravel deposits of the Cypress Hills Formation,
and precipitation of carbonate minerals at the air-rock interface, promote increasing resistance under dry conditions,
when the conglomerate is less likely to be undermined by
basal erosion or landsliding. Negative feedback also occurs
with the erosion and transport of fine material, which leaves a
lag deposit on the dominantly erosional surfaces. The
Graburn Creek watershed (Fig. I), for example, is completely
forested and the valley floor is mantled with large clasts.
Graburn Creek typically has low suspended-sediment concentrations and dilutes the sediment load of Battle Creek
(Sauchyn and Lernmen, 1996). MacPherson and Rannie
(1969) concluded that, despite 2.7 cm of floodplain accretion
and shifting of the channel during the six-day flood of May
1967,the amount of sediment exported from the basin, equivalent to basin denudation of 0.4 rnm, was less than expected of
the flood discharge.
Quasi-continuous erosion over cyclic time has also predisposed the Cypress Hills to threshold events of higher magnitude. Perhaps the best example is headward erosion and
incision of tributary valleys to a critical depth, below which
argillaceous beds are exposed and subsequently fail by sliding. The consequent exposure of poorly consolidated sediments results in greatly increased sediment production.
Landslides are the dominant mechanism of valley widening
and slope retreat, whereby the resistant caprock is ultimately
eliminated, causing a shift in type and rate of processes (i.e. a
geomorphic threshold). The impact of low-frequency events
on geomorphic history, and thus stratigraphic records, is
directly related to their magnitude and inversely related to the
passage of time. Therefore, although lowering of the plains
created the Cypress Hills over cyclic time, failure of valley
sides has had greater recent impact, especially on the
geometry of Battle Creek over graded time and the sediment
load over steady time. Relaxation times are on the order of
hundreds of years for the stabilization of valley sides and
thousands of years where landslides occupy valley floors, and
thus more directly impact the fluvial system. The observation
of multiple buried Ah horizons between slump blocks indicates that the stabilization of landslides is interrupted by episodes of renewed activity. A progressive decrease in lake
sedimentation through the late Holocene suggests decreasing
frequency and/or magnitude of landsliding, conforming to an
equilibrium model of geomorphic systems involving increasing morphological and filter resistance. Landsliding has episodically introduced large amounts of siliciclastic sediment
into the fluvial system. The deep fill in the major valleys represents a condition of disequilibrium relative to the transport
capacity of the underfit streams (Church and Slaymaker,
1989).

CONCLUSIONS
Although geologists have always utilized tectonism or climatic change to explain changes in sedimentary sequences,
or to reconstruct paleogeography, too many nagging questions remain about how the geomorphic screens function in
response to climatic change or tectonism for us to assume
that these interpretative leaps are easy or correct. The truth is
that sudden changes in a sedimentary sequence cannot be
confidently attributed to a specific geological cause until we
know more about the intermediate phase of geornorphic
response. In fact, abrupt change in the depositional sequence
is reflecting thresholds, complex response and episodic behaviour within geomorphic systems, and we are only beginning to understand how those phenomena are reflected in
sedimentary sequence. (Ritter, 1988: p. 168-169).

Thresholds, complex response, and episodic behaviour


are aspects of the geomorphic systems theory that have
emerged over the past several decades as a philosophical
framework for process geomorphology and to account for
geomorphic response to climate and tectonism at graded and
cyclic time scales. When applied to the stratigraphy and morphology of the western Cypress Hills, this theory depicts
landscape response (reaction and relaxation) to changing
boundary conditions (climate and resistance) and the inherent
instability of geomorphic systems. The reactions represent
inevitable adjustments of geomorphic systems triggered by
hydroclimatic events, but very much conditioned by geological control and geomorphic history. Quasi-continuous processes locally increase resistance to erosion, but also
predispose landscapes to processes of higher magnitude and
lower frequency and acting a different scale. The notion that
geology is an independent variable over cyclic time (Schumm
and Lichty, 1965; Hack, 1975) is contrary to the geomorphic
history of the Cypress Hills, and the stress history of the sediments underlying the glaciated northern Great Plains.
Whereas "the causes of geology, climate and initial relief'
were of "no concern" to Schumm and Lichty (1965, p. 113),
in terms of explaining the characteristics of drainage basins,
the "cause" of geology (resistance) in the Cypress Hills is
largely geomorphic. Because the surficial materials lack
inherent strength, resistance to climatic variability depends
very much on the climatic control of vegetation and pore
water, and the morphological and textural adjustment of
geomorphic systems to changing climate and extreme
weather events. A time-independent view of geomorphic process is confined to steady time and small spaces in the
Cypress Hills.
The validity of this conceptual framework, however, rests
with the stratigraphic record, the empirical basis of
geomorphic history. The existing lithostratigraphic and
biostatigraphic data from the Cypress Hills seem to support
this abstract view of the geomorphic history, but then "...the
appeal of theory is that it will produce explanations that predict phenomena of interest." (Baker and Twidale, 1991,
p. 84). A geomorphic systems perspective accents the
dependence of geomorphic response and, in turn, sedimentary sequence on scale. Because sedimentary deposits are the
product of processes operating at various spatial scales, the
complex and episodic response of slopes and channels to climatic change and hydroclimatic events should result in

D.J. Sauchyn

systematic differences among stratigraphic records that integrate geomorphic activity over various amounts of time and
space. The buried Ah horizons on valley sides indicate
repeated instability at a local spatial scale. The sediments
shed from progressively larger areas, and stored immediately
below landslides, in alluvial fans, on valley floors, and in lake
basins, should record climatic forcing of geomorphic events
of increasing magnitude and decreasing frequency.
The typical low resolution and discontinuity of stratigraphic records limits the inference of geomorphic response
to climate. Field experiments reveal the immediate responses
to hydroclimatic events, but only for current boundary conditions. Thus, intermediate spatial and temporal scales seem
most appropriate for the reconstruction of geomorphic histories with reasonable resolution, but also over sufficient time
that hydroclimatic events are superimposed on significant
changes in climate, hydrology, and vegetation. In this respect,
the emphasis of the Palliser Project on the past few millennia,
as demonstrated by the other papers in this volume, suits the
objective of establishing linkages between Holocene geomorphology and climate.

ACKNOWLEDGMENTS
Research described here was funded the Palliser Triangle
Global Change Project, the Natural Sciences and Engineering
Research Council of Canada (NSERC), the Prairie Farm
Rehabilitation Administration (PFRA), Environment
Canada, and the University of Regina. Constructive criticism
from Don Lemmen, Willem Vreeken, and an anonymous
reviewer significantly improved this paper.

REFERENCES
Ager, D.V.
1993: The Nature of the StratigraphicalRecord; Wiley, Chichester, Great
Britain, 151 p.
Bacon, C.R.
1983: Eruptive history of Mount Mazama and Crater Lake caldera, Cascade Range, U.S.A.; Journal of Volcanology and Geothermal
Research, v. 18, p. 57-1 15.
Baker, V.R. and Twidale, C.R.
1991: The reenchantment of geomorphology; Geomorphology, v. 4,
p. 73-100.
Boon, J.L.
1993: A dendrogeomorphologicalanalysis of flood frequency: Graburn
Creek, Alberta; B.Sc. thesis, University of Regina, Regina,
Saskatchewan,63 p.
Bowen, D.Q.
1978: Quaternary Geology; Pergamon Press, Oxford, United Kingdom,
221 p.
Brunsden, D.
1990: Tablets of stone: toward the ten commandments of geomo~-phology;
Zeitschrieftfiir Geomorphologie,Supplementband,v. 79, p. 1-37.
1993: Barriers to geomorphological change; in Landscape Sensitivity,
(ed.) D.S.G. Thomas and R.J. Allison; Wiley, Chichester, United
Kingdom, p. 7-12.
Brunsden, D. and T h o r n s , J.B.
1979: Landscapesensitivity and change; Institute of British Geographers,
Transactions, v. 4, p. 463484.
Bryan, R.B., Campbell, I.A., and Yair, A.
1987: Postglacial development of the Dinosaur Provincial Park badlands;
Canadian Journal of Earth Sciences, v. 24, p. 135-146.

Bull, W.B.
1991: Geomorphic Responses to Climatic Change; Oxford University
Press, New York, 326 p.
Campbell, C.
1997: Postglacial response and environmental change in southeastern
Alberta, Canada; Ph.D. thesis, University of Alberta, Edmonton,
Alberta, 187 p.
Campbell, I.A.
1982: Surface morphology and rates of change during a ten year period in
the Alberta badlands; in Badland Geolnorphologyand Piping, (ed.)
R.B.Bryan and A. Yair; Geo Books, Norwich, United Kingdom,
p. 221-238.
Carson, M.A. and Associates
1990: Off-farm sediment impacts in the SaskatchewanRiver basin; report
prepared for Water Resources Branch, Saskatchewan District,
Environment Canada, 147 p.
Christiansen, E.A. and Sauer, E.K.
1988: Age of the Frenchman Valley and associated drift south of the
Cypress Hills, Saskatchewan, Canada; Canadian Journal of Earth
Sciences, v. 25, p. 1703-1708.
Church M. and Ryder, J.M.
1972: Paraglacial sedimentation:considerationof fluvial processes conditioned by glaciation; Geological Society of America Bulletin, v. 83,
p. 3059-3072.
Church M. and Slaymaker, 0.
1989: Disequilibrium of Holocene sediment yield in glaciated British
Columbia; Nature, v. 337, p. 452-454.
Clayton, K.M.
1983: Climate, climatic change and rates of denudation; in Quaternary
Geomorphology,(ed.) D.J. Briggs and R.S. Waters; Proceedings of
the 6th British Polish Seminar, Institute of British Geographers,
Geo Books, Norwich, United Kingdom, p. 157-167.
Crickmay, C.H.
1932: The significance of the physiography of the Cypress Hills;
Canadian Field Naturalist, v. 46, p. 185-186.
1975: The hypothesis of unequal activity; inTheories of Landform Development, (ed.) W.N. Melhorn and R.C. Flemal; Allen and Unwin,
Boston, Massachusetts, p. 103-109.
Davies, G.L.H.
1989: On the nature of geo-history, with reflections on the historiography
of geomorphology; in History of Geomorphology, (ed.)
K.J. Tinkler; Binghampton Symposium No. 19, Unwin Hyman,
Boston, Massachusetts, p. 1-10.
deBoer, D.H.
1992: Hierarchies and spatial scale in process geomorphology: a review;
Geomorphology, v. 46, p. 66-77.
deBoer, D.H. and Campbell, I.A.
1989: Spatial scale dependence of sediment dynamics in a semi-arid
drainage basin; Catena, v. 16, p. 277-290.
Eybergen, F.A. and Imeson, A.C.
1989: Geomorphological processes and climatic change; Catena v. 16, p.
307-319.
Goulden, M.R. and Sauchyn, D.J.
1986: Age of rotational landsliding in the Cypress Hills, AlbertaSaskatchewan; GBographie physique et Quaternaire, v. 40,
p. 239-248.
Hack, J.T.
1975: Dynamic equilibrium and landscape evolution; in Theories of
Landscape Development, (ed.) W.N. Melhorn and R.C. Flemal;
State University of New York (Bioghampton), Publications in
Geomorphology, p. 87-102.
Janz, B. and Treffry, E.L.
1968: Southern Alberta's paralyzing snowstorms in April, 1967;
Weatherwise, April, p. 70-75, 94.
Jones, D.K.C.
1993: Global warming and geomorphology; The Geographical Journal,
V. 159, p. 124-130.
Josephson, G.J.
1988: A dendrogeomorphological investigation of a rotational landslide
in the Cypress Hills, Canada; B.Sc. thesis, Department of Geography, University of Regina, Regina, Saskatchewan, 47 p.
Jungerius, P.D.
1967: Theinfluenceof Pleistocene climatic changes in thedevelopment of
polygenetic pediments in the Cypress Hills area, Alberta;
Geographical Bulletin, v. 9, p. 218-231.

GSC Bulletin 534


Jungerius, P.D. (cont.)
1969: Soil evidence of postglacial treeline fluctuations in the Cypress
Hills area, Alberta, Canada; Arctic and Alpine Research, v. I ,
p. 235-246.
Jungerius, P.D. and Mucher, H.J.
1969: The micromorphology of fossil soils in the Cypress Hills, Alberta,
Canada; Proceedings, 3rd International Working Meeting on Soil
Micromorphology, Wroclaw, Poland, p. 617-627.
Kirkby, M.J.
1990: The landscape viewed through models; Zeitschrift fiir
Geomorphologie, Supplemenband,v. 79, p. 63-81.
Klassen, R.W.
1994: Late Wisconsinan and Holocene history of southwestern
Saskatchewan; Canadian Journal of Earth Sciences, v. 31,
p. 1822-1837.
Kulig, J.J.
1996: The glaciation of the Cypress Hills of Alberta and Saskatchewan
and its regional implications; Quaternary International, v. 32,
p. 53-77.
Kupsch. W.O. and Vonhof. J.A.
1987: .~e~ectivecementatibn
inTertiary sands and gravels, Saskatchewao;
Canadian Journal of Earth Sciences.. v. 4,- D.
. 769-776.
Last, W.M. and Sauchyn, D.J.
1993: Mineralogy and lithostratigraphy of Harris Lake, southwestern
Saskatchewan,Canada; Journal of Paleoljmnology,
-- v. 9, p. 23-39.
Leckie, D.A. and Cheel, R.J.
1989: The Cypress Hills Formation (Upper Eocene to Miocene): a
semi-arid braidplain deposit resulting from intrusive uplift;
Canadian Journal of Earth Sciences, v. 26, p. 1918-1931.
Lemmen, D.S., Dyke, L.D., and Edlund, S.A.
1993: The Geological Survey of Canada's Integrated Research and Monitoring Area (IRMA) projects: a contribution to Canadian global
change research; Journal of Paleolimnology, v. 9, p. 77-83.
Lemmen, D.S., Vance, R.E., Campbell, I.A., David, P.P.,
Pennock, D.J., Sauchyn, D.J., and Wolfe, S.A.
1998: Geomorphic systems of the Palliser Triangle, southern Canadian
Prairies: description and response to changing climate; Geological
Survey of Canada, Bulletin 521, 72 p.
MacPherson, H.J. and Rannie, W.F.
1969: Geomorphic effects of the May 1967 flood in Graburn watershed,
Cypress Hills, Alberta, Canada; Journal of Hydrology, v. 9,
p. 307-321.
Matheson, D.S. and Thomson, S.
1973: Geological implications of valley rebound; Canadian Journal of
Earth Sciences, v. 10, p. 961-978.
Middleton, G.V.
1973: Johannes Walther's law of the correlation of facies; Geological
Society of America Bulletin, v. 84, p. 979-988.
Porter, S.C., Sauchyn, D.J., and Delorme, L.D.
1999: The ostracode record from Harris Lake, southwestern
Saskatchewan: 9200 years of local environmental change; Journal
of Paleolimnology, v. 21, p. 3544.
Rhoads, B.L. and Thorn, C.E.
1993: Geomorphology as science: the role of theory; Geomorphology,
v. 6, p. 287-307.
Ritter, D.F.
1988: Landscape analysis and the search for geomorphic unity;
Geological Society of America Bulletin, v. 100, p. 160-171.
Sauchyn, D.J.
1990: A reconstruction of Holocene geomorphology and climate, western
Cypress Hills, Alberta and Saskatchewan; Canadian Journal of
Earth Sciences, v. 27, p. 1504-1510.
1993a: Quaternary and late Tertiary landscape evolution in the western
Cypress Hills; in Quaternary and Late Tertiary Landscapesof
Southwestern Saskatchewan and Adjacent Areas, (ed.)
D.J. Sauchyn; Canadian Plains Reseach Centre, University of
Regina, Regina, Saskatchewan, p. 46-58.
Sauchyn, D.J. (ed.)
1993b: Quaternary and Late Tertiary landscapes of southwestern
Saskatchewan and adjacent areas; Canadian Plains Reseach Centre,
University of Regina, Regina, Saskatchewan, 114 p.
Sauchyn, DJ. and Beaudoin, A.B.
1998: Recent environmentalchange in the southern Canadian Plains; The
Canadian Geographer, v. 42, p. 337-353.

Sauchyn, D.J. and Lemmen, D.S.


1996: Impacts of landsliding in the western Cypress H~lls,Saskatchewan
and Alberta; in Current Research 1996-B; Geological Survey of
Canada, p. 7-14.
Sauchyn, D.J. and Nelson, H.L.
1999: Origin and erosion of the Police Point landslide, Cypress Hills,
southeastern Alberta; in Holocene Climate and Environmental
Change in the Palliser Triangle: A Geoscientific Context for Evaluating the Impacts of Climate Change on the Southern Canadian
Prairies, (ed.) D.S. Lemrnen and R.E. Vance; Geological Survey of
Canada, Bulletin 534.
Sauchyn, M.A. and Sauchyn, D.J.
1991: A continuous record of Holocene pollen from Harris Lake, southwestern Saskatchewan, Canada; Palaeogeography,
Palaeoclimatology,Palaeoecology, v. 88, p. 13-23.
Sauer, E.K.
1978: The engineering significance of glacier ice-thrusting; Canadian
Geotechnical Journal, v. 15, p. 457-472.
Schumm, S.A.
1975: Episodic erosion: a modification of the geomorphic cycle; in Theories of Landscape Development, (ed.) W.N. Melhorn and
R.C. Flemal; State University of New York (Binghampton),
Publications in Geomorphology, p. 69-85.
1977: The Fluvial System; John Wiley & Sons, New York, 338 p.
1981: Erosion and response of the fluvial system - sedimentologicimplications; Society of Economic Petrologists and Mineralogists,
Special Publication No. 31, p. 19-29.
Schumm, S.A. and Lichty, R.W.
1965: Time, space and causality in geomorphology;American Journal of
Science, v. 263, p. 110-1 19.
Schumm, S.A. and Parker, R.S.
1973: Implications of complex response of drainage systems for
Quaternary alluvial stratigraphy; Nature, v. 243, p. 99-100.
Spence, C.D. and Sauchyn, D.J.
1999: Groundwater influence on valley-head geomorphology, upper
Battle Creek basin, Alberta and Saskatchewan; in Holocene
Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Stalker, A.
1965: Pleistocene ice surface, Cypress Hills area; in 15th Annual Field
Conference Guidebook, Part 1. Cypress Hills Plateau; Albcrta
Society of Petroleum Geologists, Calgay, Alberta, p. 116-130.
Thomas D.S.G. and Allison R.J.
1993: LandscapeSensitivity;Wiley, Chichester,United Kingdom,347 p.
Thorn, C.E.
1988: Theoretical Geomorphology; Unwui Hyman, Boston, Massachusetts,
247 p.
Thornes, J.B. and Brunsden, D.
1977: Geomorphology and Time; Methuen, London, United Kingdom,
208 p.
Twidale, C.R.
1993: C.H. Crickmay, a Canadian rebel; Geomorphology, v. 6,
p. 357-372.
valentine, K.W.G., King, R.H., Dormaar, J.F., Vreeken, W.J.,
Tarnocai, C., De Kimple, C.R., and Harris, S.A.
1987: Some aspects of Quaternary soils in Canada; Canadian Journal of
Soil Science, v. 67, p. 221-247.
Vance, R.E. and Last, W.M.
1994: Paleolimnology and global change on the southern Canadian Prairies; in Current Research 1994-B; Geological Survey of Canada,
p. 49-58.
Vitek, J.D. and Giardino, J.R.
1993: Preface: a perspective on getting to the frontier; in Geomorphology:
The Research Frontier and Beyond, (ed.) J.D. Vitek and J.R.
Giardino; Proceedings of the 24th Binghampton Symposium in
Geomorphology, Hamilton, Ontario, Elsevier, Amsterdam,
p. vii-xii.
Vreeken, W.J.
1986: Quaternary events at the glacial drift border near Elkwater, sootheastern Alberta; Canadian Journal of Earth Sciences, v. 23,
p. 2024-2038.

D.J. Sauchyn
Vreeken, W.J. (cont.)
1993: Loess and associated paleosols in southwestern Saskatchewan and
southern Alberta; in Quaternary and Late Tertiary Landscapes of
Southwestern Saskatchewan and Adjacent Areas, (ed.)
D.J. Sauchyn; Canadian Plains Reseach Centre, University of
Regina, Regina, Saskatchewan,p. 2745.
1999: Geomorphic surfaces and postglacial landscape evolution of the
Maple Creek basin, Saskatchewan; in Holocene Climate and
Environmental Change in the Palliser Triangle: A Geoscientific
Context for Evaluating the Impacts of Climate Change on the
Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance;
Geological Survey of Canada, Bulletin 534.
Vreeken, W.J. and Westgate, J.A.
1992: Miocene tephra beds in the Cypress Hills of Saskatchewan,Canada;
Canadian Journal of Earth Sciences, v. 29, p. 48-51.
Walker, R.G.
1984: Facies, facies sequences and facies models; in Facies Models, (ed.)
R.G. Walker; Geoscience Canada, Reprint Series 1, p. 1-9.
Walker, R.G. and James, N.P. (ed.)
1992: Facies models: response to sea level change; Geological
Association of Canada, St. John's, Newfoundland, 409 p.

Waters, P.L. and Rutter, N.W.


1984: Utilizing paleosols and volcanic ash in correlating Holocene
deposits in southern Alberta; in Correlation of Quaternary Chronologies, (ed.) W.C. Mahaney; Geobooks, Norwich, United Kingdom,
p. 203-223.
Welch, D.M.
1970: Substitution of space for time in a study of slope development;
Journal of Geology, v. 78, p. 234-239.
Westgate, J.A.
1972: The Cypress Hills; in Quaternary Geology and Geo~norphology
Between Winnipeg and the Rocky Mountains, (ed.) N.W. Rutter
and E.A. Christiansen; 24th International Geological Congress,
Field Excursion C-22, p. 50-62.
Wilson, S.E., Smol, J.P., and Sauchyn, D.J.
1997: A Holocene paleosalinity diatom record from southwestern
Saskatchewan: Harris Lake revisited; Journal of Paleolimnology,
v. 17, p. 23-31.
Wolfe, S.A., Huntley, D.J., and Ollerhead, J.
1995: Recent and late Holocene sand dune activity in southwestern
Saskatchewan; Current Research 1995-B; Geological Survey of
Canada, p. 131-140.
Wolman, M.G. and Gerson, R.
1978: Relative scales of time and effectiveness of climate in watershed
geomorphology; Earth Surface Processes, v. 3, p. 189-208.
Womack, W.R. and Schumm, S.A.
1977: Terraces of Douglas Creek, northwestern Colorado: an example of
episodic erosion; Geology, v. 5, p. 72-76.

Groundwater influence on valley-head


geomorphology, upper Battle Creek basin,
Alberta and Saskatchewan
Christopher D. spencel and David. J. sauchyn2
Spence, C.D. and Sauchyn, D.J., 1999: Groundwater influence on valley-head geomorphology,
upper Battle Creek basin, Alberta and Saskatchewan; in Holocene Climate and Environmental
Change in the Palliser Triangle: A Geoscientific Contextfor Evaluating the Impacts of Climate
Change on the Southern Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological
Survey of Canada, Bulletin 534, p. 249-255.

Abstract: Geology is the dominant control on valley morphology in the upper Battle Creek basin. At valley heads, the high infiltration capacity of the Cypress Hills Formation results in runoff being dominated by
groundwater in the form of saturation overland flow. Storm discharge events are strongly related to the antecedent moisture conditions. Exfiltrating groundwater causes seepage erosion and sapping and, in turn,
gullying. Discontinuous gullies, indicative of seepage erosion and sapping processes, occur in first-order
valleys, whereas rill development, associated with Hortonian overland flow, is apparently absent. The locations of groundwater discharge sites can be predicted with GIs and groundwater modelling, but require field
verification.
Once valleys incise to a critical depth, exposing argillaceous beds in the Ravenscrag and lower
formations, landsliding becomes the dominant process of valley evolution, masking the fluvial origin of the
tributaries. Climate is an important control on the frequency and magnitude of both mass-wasting processes
and fluvial erosion.

R6sumC : La gkologie constitue le contrdle dominant sur la morphologie des vall6es dans la partie
sup6rieure du bassin du ruissea Battle. Dans la partie amont des vallCes, la forte capacitC d'infiltration de la
Formation de Cypress Hills se traduit par le fait que 1'6coulement est domink par des eaux souterraines sous
forme de ruissellement satur6. Les Cpisodes d'kvacuation des prkcipitations sont Ctroitement lies aux conditions d'humiditk antkrieures. L'exfiltration des eaux souterraines provoque 1'6rosion par infiltration et
sapement et, ultkrieurement, par ravinement. Des ravins discontinus, qui tCmoignent de 1'6rosion par infiltration et des processus de sapement, sont pr6sents dans les vall6es de premier ordre, le d6veloppement de
rigoles associC au ruissellement hortonien Btant apparemment absent. La localisation des lieux
d'6coulement des eaux souterraines peut Ctre prCdite par la mod6lisation SIG et la modClisation des eaux
souterraines, mais elle nCcessite une vkrification sur le terrain.

partir du moment o t ~les vallks s'incisent jusqu'h une profondeur critique, exposant des couches
riches en bentonite dans la Formation de Ravenscrag et dans les formations infkrieures, les glissements de
terrain deviennent le processus dominant d'bvolution des vallBes, ce qui masque l'origine fluviale des
tributaires. Le climat constitue un contr8le important de la frCquence et de la magnitude des processus de
mouvement en masse et de 1'Qosion fluviale.

' Atmospheric and Hydrologic Sciences Division, Environment Canada, 301-5204 50th Avenue, Yellowknife,
Northwest Territories X I A 1E2
Department of Geography, University of Regina, Regina, Saskatchewan S4S 0A2

GSC Bulletin 534

INTRODUCTION
The slope geomorphology of major valleys in the Cypress
Hills, and elsewhere in the Palliser Triangle, is dominated by
landslide deposits and reflects the dominance of
mass-wasting processes (Sauchyn, 1993, 1999). Nonetheless, it is the exposure of argillaceous bedrock by fluvial erosion that makes these slopes prone to f a i l u r e . ~ e s ~ ithis
te
fundamental causal relationship, processes of fluvial erosion
in this region remain poorly investigated. This study contributes to our understanding of these processes by examining the
role of groundwater discharge as a control on valley-head fluvial process in the upper Battle Creek watershed (Fig. 1).
Groundwater discharge in the Cypress Hills is controlled
by either geological contacts or topography. Discharge
occurs along geological contacts where a highly permeable
formation (aquifer) overlies a formation of significantly
lower permeability (aquitard), such that groundwater will
flow along the contact. This type of groundwater discharge is
commonly associated with landsliding. In contrast, topographic discharge of shallow groundwater is controlled by
changes in topographic slope and surficial geology. This type
of groundwater discharge has the potential to directly influence valley-head fluvial erosion, and is therefore emphasized
in this studv.
Hortonian runoff ('infiltration excess overland flow';
Pearce et al. (1986)) is uncommon in the Cypress Hills,
because the infiltration capacity of most soils exceeds rainfall
intensity. Permeable surfaces and low rainfall intensity result
in water infiltration and flow via subsurface pathways. At the
limit of subsurface water storage and transmission, gravity
water is discharged to the surface to produce 'saturation overland flow' (Dunne, 1990). The emerging groundwater erodes
by seepage erosion and sapping (Higgins, 1984; Dunne,
1990; Moeyersons, 1991). Shear stress on the margins of
pores entrains soil particles, and may cause sapping when
overlying soil collapses, creating gullies (Baker, 1990).

This study attempts to evaluate the influence of groundwater discharge on valley-head erosional processes, through
examination of the hydrology and geomorphology of tributaries to upper Battle Creek. Slow and moderate changes in
stream flow and dissolved solids are hydrological indications
of groundwater storm flow in small basins. Discontinuous
gullies emanating from valley heads are considered morphological evidence of seepage erosion and sapping, and therefore were a focus of field mapping. In order to assist field
mapping of groundwater discharge sites, a steady-state model
was applied within a geographic information system (CIS) to
predict the locations of both contact and topographic springs
in the Battle Creek basin.

STUDY AREA
The Cypress Hills, rising up to 600 m above the adjacent
plains, are an erosional remnant of Late Tertiary landscapes
(Russell and Wickenden, 1933). The west block (Fig. 1) is the
highest of a series of dissected plateaus, with its nonglaciated
surface having formed a 300 km2 nunatak at the maximum of
the last glaciation (Stalker, 1965). Meltwater drainage
through the west block incised major preglacial valleys and
caused tributary streams to degrade to a lower base level.
The lower slopes of the Cypress Hills are underlain by the
Upper Cretaceous Bearpaw Formation (Table I), predominantly marine shales with numerous bentonitic beds
(Furnival, 1946). The overlying Eastend, Whitemud, Battle,
and Frenchman formations include sandstone, mudstone,
shale, clay, bentonite, and ironstone. Tertiary sediments consist of the Ravenscrag Formation, composed of thinly bedded
sandstone with lignite interbeds, and the capping Cypress
Hills Formation, composed of sand, gravel, and conglomerate (Leckie and Cheel, 1989). Almost all slopes have been
subject to landsliding, principally due to failure of clay-rich
beds in the Ravenscrag, Frenchman, and Battle formations.

11030'W
49'45' N

Alberta

police ~ d n t

Saskatchewan

Figure I .
Location of the Cypress Hills and upper Battle
Creek basin in southeastern Alberta and southwestern Saskatchewan. Cypress Hills plateau is
shaded, lakes are black. Location of the Cypress
Hills within the Palliser Triangle is shown on the
small-scale inset map.

C.D. Spence and D.J. Sauchyn

Table 1. Bedrock geology and associated hydraulic


conductivities, Cypress Hills (Furnival, 1946; Freeze
and Cherry, 1979; Meneley, 1983; Weeks and
Gutentag, 1988; Lennox et al., 1988; Leckie and Cheel,
1989).

Thickness
(m)

Epoch
Oligocene
Paleocene
Upper Cretaceous
Upper Cretaceous
Upper Cretaceous
Upper Cretaceous
Upper Cretaceous

Cypress Hills
Ravenscrag
Frenchman
Battle
Whitemud
Eastend
Bearpaw

I1

Hydraulic
conductivity
(mdl)

Spring mapping and sampling

15-76
70+
345+
6-9
10-14
21-37
285-305

Field mapping along valley slopes and bottoms was conducted to locate and characterize groundwater discharge, and
to test the value of CIS results. Springs were characterized
(Table 2) by surface expression, location and, where flow was
sufficient, sulphate content, which is used as a proxy measure
of subsurface residence time (Freeze and Cherry, 1979).
Spring water was filtered through a 63 prn mesh to remove
sediment, after which barium chloride was added to form barium sulphate. Light absorption by the barium sulphate in suspension was measured with a photometer and the sulphate
concentration determined by reading a standard curve. Accuracy of this method is expected to be within 10%.

Runoff hydrology
Such failures occur during periods of excess groundwater and
high pore-water pressure, which reduce the shear strength in
the clay beds (Sauchyn and Nelson, 1999).
The climate of the Cypress Hills is subhumid and cool.
Mean annual temperature and precipitation are 2.5"C and
457 mm (Cowell, 1982). Approximately 70% of annual precipitation occurs in May and June. Local climates vary considerably between the forested slopes of the valleys and the
open grassland plateaus.

METHODS
GZS modelling
Benson, Graburn, and Fort Walsh creeks, all tributaries to
upper Battle Creek, were analyzed using ARCONFO GIs to
predict the location of both topographic and contact groundwater discharges. The GIs parameters included slope,
surficial geology, hydrography, and topography.
Topographic discharge of groundwater was predicted
using a form of O'Loughlin's (1986) flow model:

where A is the area of the map unit, M is the slope, b is the


contour interval (5 m), and K is the hydraulic conductivity of
the surficial geological unit. The product of A, M, and K is q.
The probability of contact discharge is a function of the difference between the conductivities of the two formations
(Freeze and Cherry, 1979). For this study, differences of
greater than lOOx were used to define potential areas for contact springs (e.g, at the base of the Cypress Hills and Eastend
formations, Table 1). This model predicts only the distribution, and not the discharge, of springs. Mapping the magnitude of groundwater discharge would require the evaluation
of hydroclimatic parameters, and a seasonal (or higher resolution) temporal component to the model.

The hydrology and water chemistry of a 1 km2 first-order


stream valley in theFort Walsh Creek basin (Fig. 1) was monitored to determine the significance of groundwater discharge
during storms. A V-notch weir was used to channel
streamflow into a bucket where flow could be measured volumetrically. Rainfall was measured at two unobstructed sites.
The hillslopes were observed throughout the runoff event for
saturation overland flow and infiltration excess overland
flow. Infiltration was measured a few hours after rainfall
began, to determine the infiltration capacity of the soils
during storms.
Conductivity of rainwater and stream-water samples was
determined to evaluate the contribution of groundwater to
storm runoff. Samples were collected at two-hour intervals
and analyzed using a conductivity probe, with an expected
accuracy of 1%.

Morphometry
Long profiles of three, first-order valley heads were surveyed
using a 50 m tape measure, Abney level, and stadia rods. Particular attention was paid to the occurrence of discontinuous
gullies, considered diagnostic of seepage erosion and sapping
(Patton and Schumm, 1975).

RESULTS
GZS modelling
Predicted occurrences of groundwater discharge were similar
for all three basins modelled. Within Benson Creek, discharge was predicted to occur at the mouth of the basin, at the
tributary valley bottoms with decreased topographic slope
(concavity) and hydraulic conductivity, and along the contacts between the Cypress Hills and Ravenscrag formations
and the Eastend and Bearpaw formations (Fig. 2). Field mapping revealed four springs in the Benson Creek basin, two
downstream of the predicted locations in the upper basin and
two within the predicted locations near the mouth. This suggests that modelling was reasonably successful in predicting
discharge locations.

GSC Bulletin 534

Table 2. Summary of spring characteristics.

Valley order
Second
First
Third
Second
Second
Third
Second
Sixth
First
Second
Second
Second
Second
Second
Second
Second
Second
Second
Third
Third
Sixth
Seventh
Seventh

Sulphate
(mg.L1)
0.0
0.0
0.0
2
6.0
8.0
9.6
16.5
0.0
5.8
7.0
7.0
7.0
12
0.0
0.0
1.O
2
6.4
8.5
15.7
22.5
50.0

Formation of
origin
Ravenscrag
Cypress Hills
Ravenscrag
Whitemud
Cypress Hills
Ravenscrag
Cypress Hills
Whitemud
Cypress Hills
Cypress Hills
Cypress Hills
Cypress Hills
Cypress Hills
Cypress Hills
Cypress Hills
Ravenscrag
Ravenscrag
Whitemud
Whitemud
Ravenscrag
Ravenscrag
Whitemud
Bearpaw

Surface
expression'
Gully
Gully
Gully
Gully
Gully
Gully
Gully
Gully
Rocky
Rocky
Lobate
Lobate
Lobate
Lobate
Marsh
Marsh
Marsh
Marsh
Marsh
Marsh
Marsh
Marsh
Marsh

Spring location
Valley bottom
Valley bottom
Valley bottom
Valley bottom
Valley bottom
Valley bottom
Valley bottom
Valley bottom
Valley bottom
Valley side
Valley side
Valley side
Valley side
Valley side
Valley bottom
Valley bottom
Valley bottom
Valley bottom
Valley bottom
Valley bottom
Valley side
Valley side
Valley side

Discharge
agent
Topographic
Topographic
Topographic
Topographic
Indeterminate
Topographic
Indeterminate
Topographic
Topographic
Contact
Contact
Contact
Contact
Contact
Topographic
Topographic
Topographic
Topographic
Topographic
Topographic
Contact
Contact
Contact

Basin
Fort Walsh
Battle
Benson
Fort Walsh
Battle
Benson
Graburn
Nine Mile
Battle
Storm
Graburn
Battle
Battle
Graburn
Graburn
Ranger
Fort Walsh
Fort Walsh
Fort Walsh
Ranger
Nine Mile
Battle
Battle

'Four distinct surface expressions of groundwater discharge sites were observed:


Lobate springs are located on valley sides, and identified by a phreatophyte-covered lobe of regolith that
overhangs the discharge point. The lobe morphology is the product of slow mass wasting of saturated
surface materials.
Marsh springs, located in flatter sites, reflect dispersed discharge owing to local morphological factors.
Gully springs form on slopes by seepage erosion and sapping.
Rocky springs have a gully morphology but the discharge point is dominated by boulders, the result of
spring water having washed away the fine sediment.

Predicted topographic discharge sites


Predicted contact discharge sites

C.D. Spence and D.J. Sauchyn

Figure 3.
Rainfall, streamflow, and conductivity in Fort
Walsh Creek basin during storm, July 9 and 10,
1992.

Time (h)

Springs

Runoff hydrology

Although 120 springs were located through field mapping,


only 23 had sufficient flow to allow water sampling. These
springs are listed in Table 2, grouped by surface expression
and sulphate content. The concentration of sulphate should
increase as groundwater moves down from the plateau
through the underlying shale (Mackay et al., 1936;
Rutherford, 1967). Of the 18 springs with low (<2 m g . ~ - l )
sulphate content, 14 were located at higher elevations in firstand second-order valleys. The four discharges with sulphate
concentrations greater than 15 m g . ~ - lall occurred within
sixth- and seventh-order valleys. Therefore, sulphate concentration does appear to be a reliable indicator of subsurface residence time of groundwater in the Battle Creek basin.

Detailed data are available for one storm and subsequent runoff at a first-order valley head in the Fort Walsh Creek basin
(Fig. 3). Precipitation fell from 02:30 h on July 9, 1992 to
19:30 h on July 10, and totalled 38.5 mm. Antecedent rainfall
was 70 mm in the previous week, but no significant increase
in flow had been observed at the valley head. The rainfall
peak, accounting for 30% of the total rainfall, occurred within
the first four hours of the storm (Fig. 3). Rainfall intensity was
also highest during this period (6 rnm.h-I). Rainfall during the
first hour was 8% of the weekly total, whereas streamflow
during this same period more than doubled from 700 cm3.s-'
to 1600 cm3.s-l. After another 6 mm of rainfall in the next
hour, the flow doubled again to 2800 cm3.s-l. Stream discharge peaked 34 h after the start of the storm at 8500 cm3.s-l,
or 7750 cm3.s-l above base flow (Fig. 3). Conductivity of the
stream water decreased from 510 @.cm-I at the start of the
storm to a minimum of 400 @.cm-l, and then increased after
the runoff event to 520 M.cm-l (Fig. 3). The conductivity at
peak flow was 440 @.cm-l, whereas the conductivity of the
rainwater averaged 7 @.cm-l. No Hortonian runoff was
observed during the runoff event.

Sulphate concentration, along with surface expression


and spring location, were considered when determining the
discharge agent of a spring (contact versus topographic
springs). Sulphate concentration is the most important characteristic within first- and second-order valleys. A combination of gully, rocky, or marshy morphology, valley-bottom
location, and low sulphate suggest the spring is topographically induced. In contrast, a combination of lobate morphology, valley-side location, and high sulphate concentration
suggest the spring is controlled by geological contacts. Other
combinations were considered indeterminate. In higher order
valleys, where all of the groundwater has had a long residence
time and thus a high sulphate concentration, only the surface
expression and the location of the spring are used to
determine the discharge agent (Table 2).

Figure 2.
Results of the GIs modelling of groundwater discharge sites,
Benson Creek. Heavy black line on topographic surface
denotes boundary of drainage basin.

The runoff and conductivity data from this valley-head


site indicate that saturation overland flow was the dominant
runoff process. The 34 h lag time for a 1 km2 basin is a strong
indication that exfiltrating groundwater is the major contributor to stormflow (Newbury et al., 1969; Sklash et al., 1986).
The disproportionate increases in the volume of streamflow
compared to total rainfall during the first few hours of the
storm also suggest contributions from a source other than
rainfall, that is, antecedent water. Finally, the fact that stream
water maintained a relatively constant conductivity throughout the runoff event, at levels far above that of rainwater, is
only possible where saturation overland flow of pre-storm
water dominates. The steady rise in conductivity during flow
recession also supports this conclusion.

GSC Bulletin 534

Figure 4,

100

200

300

400

500

600

700

Horizontal distance (m)

Figure 5.
Gullying at the confluence of twofirst-order valley heads. The concentration of subsut$acefZow
paths at the mouths of the valleys causes groundwaterflows to exceed soil transmissivity, leading
to exfiltration and the creation of the gullies.
Photograph by C.D Spence. GSC 1999-039

Morphometric analysis
If saturation overland flow is indeed the dominant runoff process at valley heads, as demonstrated above by the July 9
event, these sites should show morphological evidence of
seepage erosion and sapping. Discontinuous gullies, evident
as abrupt steps along the longitudinal profiles of first-order
valley heads (Fig. 4), are interpreted to be the product of these
processes. In contrast, rills, which are characteristic of
Hortonian flow, were not observed in the valley heads
(Fig. 5).
This high infiltration capacity of the sand and gravel units
of the Cypress Hills Formation prevents Hortonian flow in
most first-order valleys. Field tests during this study estimated these rates to be about 6 cm.min-l. Although the accuracy of these tests is limited, giving results two to ten times
the actual capacity (cf. Dunne and Leopold, 1978), even 10%
of the measured infiltration capacity would be sufficient for
most rainfall to infiltrate.
Whereas morphological evidence exists for groundwater
discharge and erosion at the valley heads, the fluvial origin of
the tributary valleys at a basin scale is disguised by mass
wasting. The fluvial morphology of valley heads ends
abruptly at a critical threshold of valley depth, incision
beyond which exposes the shales of the Ravenscrag
Formation, and virtually all valley sides assume the morphology characteristic of landslides (Sauchyn and Goulden,
1988).

DISCUSSION
Groundwater is an effective geomorphic agent in the Cypress
Hills of southwestern Saskatchewan and southeastern
Alberta. Although landslides, caused by shearing of clay-rich
beds under excess pressure, dominate the slope geomorphology, nearly all valleys have originated from headward growth
of tributary valleys. Discontinuous gullies, indicative of
seepage erosion and sapping associated with the exfiltration
of groundwater (Higgins, 1984; Dunne, 1990; Moeyersons,
1991), are the dominant feature of fluvial erosion at valley
heads. Hortonian overland flow and associated rill development are absent.
By successfully predicting the location of groundwater
discharge, the GIs model also serves to predict those slopes
most prone to gullying and instability. However, attempts to
model the dynamics of seepage erosion, sapping, and
landsliding processes in the Cypress Hills must incorporate
key climatic and hydrogeological parameters, and be complemented by monitoring studies. Monitoring of the groundwater table has been initiated as part of a study of the
instability of the Police Point landslide (Sauchyn and Nelson,
1999).
Although geology is the dominant control on valley evolution in the Cypress Hills, climate is an important factor controlling the frequency and magnitude of both fluvial erosion
and mass-wasting processes (e.g. Sauchyn and Lemmen,
1996). For example, in the storm documented in this study,

C.D. Spence and D.J. Sauchyn

76 mm of rainfall during the previous week was required to


initiate an increase in first-order stream flow. A shift towards
a drier climate would generally enhance slope stability,
although extreme hydrological events will continue to induce
significant geomorphic response regardless of shifts in mean
climate conditions (e.g. Sauchyn, 1999).

ACKNOWLEDGMENTS
Funding for this research was provided by a Palliser Triangle
Global Change Research Grant, the Saskatchewan Wheat
Pool, Alice M. Goodfellow, and Clifton Associates Ltd. The
authors wish to thank Gregg Ambrosi, Iain Stewart, IOm
Hodge, Troy Riche, Robert Stevenson, and Cheryl Senft, who
assisted with fieldwork. This manuscript benefited greatly
from the comments of reviewers W.F. Rannie and V. Sloan.
Our thanks also go to Don Lemmen for his input into the final
form of the manuscript.

REFERENCES
Baker, V.R.
1990: Spring sapping and valley network development; in Groundwater
Geomorphology, (ed.) C.G. Higgins and D.R. Coates; Geological
Society of America, Boulder, Colorado, p. 235-265.
Cowell, W.
1982: Cypress Hills Provincial Park climate; Alberta Recreation Parks
and Wildlife, Edmonton, Alberta, 11 p.
Dume, T.
1990: Hydrology, mechanics, and geomorphic implications of erosion by
subsurface flow; in Groundwater Geomorphology, (ed.)
C.G. Higgins and D.R. Coates; Geological Society of America,
Boulder, Colorado, p. 1-28.
Dunne, T. and Leopold, T.B.
1978: Water in Environmental Planning; W.H. Freeman and Company,
New York, 818 p.
Freeze, R.A. and Cherry, J.A.
1979: Groundwater;Prentice Hall, Englewood Cliffs, New Jersey, 604 p.
Furnival, G.M.
1946: Cypress Lake map area, Saskatchewan; Geological Survey of
Canada, Memoir 242, 161 p.
Higgins, C.G.
1984: Piping and sapping: developmentof landformsby groundwateroutflow; in Groundwater as a Geomorphic Agent, (ed.) R.A. LaFleur;
Allen and Unwin, Boston, Massachusetts,p. 18-58.
Leckie, D.A. and Cheel, R.J.
1989: The Cypress Hills Formation (Upper Eocene to Miocene): a
semi-arid braidplain deposit resulting from intrusive uplift;
Canadian Journal of Earth Sciences, v. 26, p. 1918-1931.
Lennox, D.H., Maathuis, H., and Pederson, D.
1988: Western glaciated plains; in Hydrogeology, (ed.) W .Back,
J.S. Rosenshein, and P.R. Seaber; Geological Society of America;
Geology of North America, Boulder, Colorado, v. 2, p. 115-128.
Mackay, B.R., Beach, H.H., and Goodall, D.P.
1936: Groundwater resources of the Rural Municipality of Cypress Hills
#80, Saskatchewan; Geological Survey of Canada, Water Supply
Paper 224,44 p.

Meneley, W.A.
1983: Hydrogeology of the Eastend to Ravenscrag Formations in southern
Saskatchewan; Saskatchewan Environment, Regina,
Saskatchewan, 21 p.
Moeyersons, J.
1991: Ravine formation on steep slopes: forward versus regressive
erosion - some case studies from Rwanda; Catena, v. 18,
p. 309-324.
Newbury, R.W., Cherry, J.A., and Cox, R.A.
1969: Groundwater - streamflow systems in Wilson Creek Experimental
Watershed, Manitoba; Canadian Journal of Earth Sciences, v. 6,
p. 613-623.
O'LoughUn, E.M.
1986: Prediction of surface saturation zones in natural catchments by topographic analysis; Water Resources Research, v. 22, p. 794-804.
Patton, P.C. and Schumm, S.A.
1975: Gully erosion, northwestern Colorado: a threshold phenomena;
Geology, v. 3, p. 88-90.
Pearce, A.J., Stewart, M.K., and Sklash, M.G.
1986: Storm runoff generation in humid headwater catchments: 1. Where
does the water come from? Water Resources Research, v. 22,
p. 1263-1274.
Russell, L.S. and Wickenden, R.T.D.
1933: An upper Eocene vertebrate fauna from Saskatchewan; Transactions of the Royal Society of Canada, ser. 3, v. 27, p. 35-65.
Rutherford, A.
1967: Water quality survey of Saskatchewan groundwaters;
Saskatchewan Research Council, Report C-66-1.285 p.
Sauchvn.
" ,D..T
-.
1993: Quaternary and late Tertiary landscape evolution in the western
Cypress Hills; in Quaternary and Late Tertiary Landscapes of
Southwestern Saskatchewan and Adjacent Areas, (ed.)
D.J. Sauchyn; Canadian Plains Research Center, Regina,
Saskatchewan, p. 46-58.
1999: Geomorphology of the western Cypress Hills: climate, process,
stratigraphy and theory; in Holocene Climate and Environmental
Change in the Palliser Triangle: A Geoscientific Context for
Evaluating the Impacls of Climate Change on the Southern
Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological
Survey of Canada, Bulletin 534.
Sauchyn, D.J. and Goulden, M.R.
1988: The role of landsliding in the evolution of the Cypress Hills; Regina
Geographical Studies #5, (ed.) A. Paul and R. Widdis; Department
of Geography, University of Regina, Regina, Saskatchewan,
p. 63-80.
Sauchyn, D.J. and Lemmen, D.S.
1996: Impacts of landsliding in the western Cypress Hills, Saskatchewan
and Alberta; in Current Research 1996-B; Geological Survey of
Canada, p. 7-14.
Sauchyn, D.J. and Nelson, H.L.
1999: Origin and erosion of the Police Point landslide; in Holocene
Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Context for Evaluating the Impacts of Climate
Change on the southern Canadian Prairies, (ed.) D.S. Lemrnen and
R.E. Vance; Geological Survey of Canada, Bulletin 534.
Sklash, M.G., Stewart, M.K., and Pierce, A.J.
1986: Storm runoff generation in humid headwater catchments: 2. A case
study of hillslope and low order stream response; Water Resources
Research, v. 22, p. 1273-1287.
Stalker. A.
1965: Pleistoceneice surface, Cypress Hills area; in Cypress Hills Plateau,
Alberta and Saskatchewan. 15IhAnnual Field Conference Guidebook, Part 1, (ed.) R.L. ell; Alberta Society of Petroleum Geologists, Calgary, Alberta, p. 116-130.
Weeks, J.B. and Gutentag, F.D.
1988: High plains; in Hydrogeology, (ed.) W .Back, J.S. Rosenshein, and
P.R. Seaber; Geological Society of America, Geology of North
America, v. 2, p. 157-164.

Origin and erosion of the Police Point landslide,


Cypress Hills, Alberta
David J. sauchynl and Hugh L. el son^
Sauchyn, D. J, and Nelson, H.L., 1999: Origin and erosion of the Police Point landslide, Cypress
Hills, Alberta; in Holocene Climate and Environmental Change in the Palliser Triangle: A
Geoscientific Contextfor Evaluating the Impacts of Climate Change on the Southern Canadian
Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological Survey of Canada, Bulletin 534,
p. 257-265.

Abstract: The Police Point landslide is the largest historic landslide in the Cypress Hills, and is typical of
the complex slope failures that dominate the slope geomorphology of the Cypress Hills and major valleys of
the Palliser Triangle. The failure produced proximal slump blocks and distal earth flows which continue to
slowly creep and slide. Continuous sediment production by rill erosion, mass wasting of secondary scarps,
and gully erosion of the earthflow deposits impacts forest, riparian, and aquatic ecosystems. Average net
erosion of 6.4 cm at 101 steel rods, measured from October 27,1994 to July 3,1996, represents 2290 m3 of
sediment loss from about 0.35 krn2 of landslide deposits. Up to 45 cm of gully erosion occurred during single storms. Significant variability among pins and observation days reflects the threshold response of surface sediments to hydrological and meteorological conditions, whereby much of the annual sediment
redistribution occurs during a few runoff events.
R6sumC : Le glissement de la pointe Police, principal glissement de terrain de la pBriode historique dans
les collines Cypress, est reprtsentatif des ruptures de pente complexes qui dorninent la gtomorphologie des
versants des collines Cypress et des principales vall6es du Triangle de Palliser. Ce glissement a engendre
des blocs de dkcrochement proximaux et des coultes de terre distales, dont les mouvements de reptation et
de glissement se poursuivent lentement. La production continue de sCdiments par Crosion des rigoles,
mouvements en masse des escarpements secondaires et trosion en ravins des dCp8ts des coulCes de terre se
rkpercutent sur les Ccosystkmes forestiers, riverains et aquatiques. Le taux d'Crosion moyen net de 6,4 cm
mesurt ?I 101 tiges d'acier du 27 octobre 1994au 3 juillet 1996reprCsenteune perte de 2290 m3de stdiments
B partir de 0,35 km2 de dCp8ts de glissement. L'Brosion par ravinement B la suite de chutes de pluie
individuelles atteint une valeur maximale de 45 cm. La variabilitC importante entre pieux et entre jours
d'observation reflbte le seuil de rCaction des sCdiments de surface aux conditions hydrologiques et
mCtBorologiques, le dkplacement annuel des skdiments se produisant en bonne partie 9 l'occasion de
quelques Bpisodes d'kcoulement.

Department of Geography, University of Regina, Regina, Saskatchewan S4S 0A2


Environment Canada, 2365 Albert Street, Regina, Saskatchewan S4P 4K1

GSC Bulletin 534

INTRODUCTION
In the Palliser Triangle, landslides are "ubiquitous"
(Thomson and Morgenstern, 1978, p. 516), and are the dorninant process of valley widening in large meltwater valleys
and along deeply incised tributaries (de Lugt and Campbell,
1992). For example, the Frenchman River valley, south of the
eastern Cypress Hills (Fig. I), is 80 m less deep and almost
three times wider than the original meltwater channel, with
the basal valley fill composed mainly of landslide debris
(Christiansen and Sauer, 1988). The most recent major slope
failure in the western Cypress Hills occurred in mid-May
1967, when an estimated 1.5 x lo6 m3 of Tertiary and Upper
Cretaceous bedrock moved down the south side of Battle
Creek valley near Police Point, about 3 km west of the
Alberta-Saskatchewan boundary (Mark Engineering
Geology Ltd., unpub, report for Alberta Recreation and
Parks, 1967, Fig. 1).

This paper provides a brief overview of the causes of


landsliding in the Palliser Triangle, followed by a more
detailed description of the origin and subsequent erosion of
the Police Point landslide. Sediment shed from the landslide
accumulates on the lower valley side and enters Battle Creek,
where it alters stream geometry and water quality, affecting
fish productivity and reproduction by inundating food supplies, eggs, and spawning beds (Sauchyn and Lemmen,
1996). Because land use in the vicinity of the landslide is limited to cattle grazing and recreation, the failure of the slope
itself had far less ecological and human impact than has the
subsequent runoff and erosion. Therefore, emphasis in this
paper is placed on the nature and rate of erosion of the
landslide.

Figure 1. Location of the western Cypress Hills and the Police Point landslide. On
upper map, dashed line denotes limit of Palliser Triangle, light shaded area is Brown
Chernozemic Soil Zone, and dark shading indicates the three "blocks" of the Cypress
Hills. On lower map, shading indicates upper plateau surface (Cypress Plain) and
arrow indicates location and direction of landslide.

D.J. Sauchyn and H.L. Nelson

CAUSES OF LANDSLIDING
The overriding cause of landsliding in the Palliser Triangle is
the low shear strength of the Upper Cretaceous shale, especially those containing bentonite (reviewed by Sauchyn in
Lemmen et al. (1998)). Other important geological and
geomorphic factors promoting landsliding include
1) overconsolidation and unloading of the Cretaceous shale
from the weight of Tertiary sediments and Pleistocene ice
sheets, 2) local deformation from ice thrusting and rebound of
strata under incising valleys, 3) rapid incision by glacial meltwater and postglacial streams, 4) regional fracturing of bedrock and Quaternary sediments, 5) perched groundwater
tables, and 6) lateral shifting of stream channels (Mallard,
1977; Thomson and Morgenstern, 1977,1978).
In the western Cypress Hills, Tertiary sand, gravel, conglomerate, and fine-grained sandstone overlie Upper
Cretaceous shale, clay, silt, and sand (Table 1). In the absence
of glacial drift, rainwater and snowmelt water readily infiltrate the coarse permeable caprock and seep to the
fine-textured bedrock of low hydraulic conductivity and
shear strength. It is estimated that more than 90% of the
"clay-like sediments" in the Cypress Hills are the high montmorillonite (bentonitic) shale of the Upper Cretaceous
(Lindoe, 1965,p. 210). This geological setting, the relatively
steep slopes,
local relief of up to 200 m, and about 100 mm
more precipitation than on the surrounding plains have
together resulted in deep-seated landsliding and the terraced
and hummocky topography that dominate the slope geomorphology of the Cypress Hills (Fig. 2; Sauchyn, 1993).

conductivities (10.~to 10-~cm.s-'; McCann & Associated


Ltd., unpub. report for David Brornley Engineering (1983)
Ltd., 1987). In the clay beds, volumetric water contents are
consistently above plastic limits, with plasticity indices of
20-30% (medium-high plasticity). Two potential slope failure mechanisms are recognized: shallow rotational failure of
weathered bedrock and deep-seated block failure of highly
plastic clay beds (EBA Engineering Consultants Ltd., unpub.
report for Alberta Environmental Protection, 1996). The
most common are high-frequency-low-magnitude failures
that produce short scarps (0.2-0.6 m) on the upper parts of
hillsides. Deep-seated failures, including the Police Point
landslide, require saturated conditions and excessive
pore-water pressure in one or more beds of plastic clay, and
are far less probable despite a low factor of safety (shear
strength / shear stress < 1.25) for most slopes.
Table 1. Table of formations, west block, Cypress Hills
(after Furnival, 1946; Leckie and Cheel, 1989).

EoceneMiocene
Paleocene

The engineering properties of the Upper Cretaceous bedrock are known from a number of geotechnical evaluations in
the western Cypress Hills. Because nearly all slopes have
been disturbed by landsliding, there is no consistent borehole
stratigraphy. Landslides are related to shallow (<1 m)
perched groundwater systems and to weathered and/or previously mass-wasted Cretaceous bedrock with low hydraulic

Upper
Cretaceous

Ravenscrag

70+

Frenchman

3-45+

Battle
Whitemud

6-g

10-14

Eastend

21-37

Bearpaw

285-305

sandstone, sand,
gravel
Sand, silt, clay,
lignite
Sandstone, shale

Bentonitic shale,
silt, sand
Clay, sandstone,
silt
Sand, silt, clay,
lignite
Marine shale,
sandstone

Figure 2.
Dormant late Holocene landslide on the north
side of the Battle Creek valley about 1 km west
of the Alberta-Saskatchewan boundary.
GSC 1999-040A

GSC Bulletin 534

Vertical aerial photographs document the evolution of the


Police Point area before and after the failure of 1967. In 1952
(Fig. 4a), there was no evidence of slope instability. An active
section of plateau rim appears in the 1962 airphoto (Fig. 4b),
but the valley side still has continuous vegetation cover. The
first large-scale airphoto of the 1967 landslide, taken in 1969
(Fig. 4c), shows that the vertical scarp which had begun forming by 1962 had extended westward. A section of Cypress
Plain dropped along the fracture visible in 1962, while catastrophic mass wasting and earthflows along the east side of
the slide destroyed the forest cover. The survival of forest on
the west side suggests deep-seated sliding and lack of flowing: trees were tilted but not removed. By 1981 (Fig. 4d), erosion and shallow mass wasting had significantly reduced the
tree cover. The most recent large-scale airphoto, from 1992
(Fig. 4e), shows that relatively few changes occurred on the
landslide during the predominantly dry years of the 1980s.

EROSION OF THE LANDSLIDE

Figure 3. Vertical aerial photograph of the Police Point


landslide from 1981 showing 1) the Cypress Plain, 2 ) the
main scarp, 3)proximal slump blocks and secondary scarps,
4) the distal slump block monitoredfor erosion (see Fig. 6),5)
the earthflow deposit which constitutes much of the lower
part of the landslide, and 6)gullies incised into the earthflow.
Bottom of landslide lies about 140 m below the elevation of
the Cypress Plain. Photograph courtesy of Alberta
Environmental Protection, Air Photo Services (photo no.
BR83059 L1 OE-124).

ORIGIN OF THE POLICE POINT LANDSLIDE


The Police Point landslide occurred on or shortly after May
14, 1967, following a dramatic rise in air temperature after
several weeks of cool weather and 1.5 m of snowfall (Janz
and Treffry, 1968). The magnitude of the triggering
snowmelt event was reflected by flooding of nearby Graburn
Creek (Fig. I), which had a return period of at least 50 years
(MacPherson and Rannie, 1969). Although there are no
first-hand accounts of the slope failure, excessive pore water
would have prompted the shearing of clay(s), producing a
typical composite slope failure, with proximal rotational
slumping and distal earthflows (Fig. 3). Kent and Simpson
(1973) concluded that the proximal slump blocks moved
along fractures caused by the failure of bentonite and
bentonitic clay at the base of the Cypress Hills Formation.
Downslope, mass wasting would have required the failure of
lower beds of plastic clay, most likely in the Battle and
Eastend formations, which are rich in bentonite (Lindoe,
1965).

Since 1967, the Police Point landslide has been extensively


modified by mass wasting and erosion of the scarps and
headward erosion of gullies. The main scarp (Fig. 3) retreats
as the matrix of sand is washed out and large clasts fall to the
base of the scarp (Fig. 5a). Steep proximal secondary scarps
(Fig. 3) in fine sediments retreat by rill erosion (Fig. 5b). A
steep distal scarp (Fig. 3) has failed by debris flow and shallow sliding (Fig. 5c). The earthflow deposits that comprise
the lower part of the landslide (Fig. 3) are subject to severe
gully erosion (Fig. 5d).
Given this morphological evidence of geomorphic activity, the lack of vegetation after nearly 30 years, and the
downslope and downstream impacts of the sediments produced on the landslide, field research was initiated in 1994 to
examine the nature and rate of erosion of the landslide deposit
and delivery of suspended sediments to Battle Creek
(Sauchyn and Lemmen, 1996). The first priority was the
monitoring of erosion. On October 27,1994,101 erosion pins
(steel rods) were installed along eight transects (tiers) on the
earthflow deposits and along two transects on a distal secondary scarp (Fig. 6). A detailed topographic (total station) survey of the landslide in 1995 yielded 867 elevations, including
the co-ordinates of the erosion pins and three control points.
A subsequent survey will provide data on any displacement
of the erosion pins due to mass movement of the sediments.
The landslide was mapped again in 1996, using differential
GPS. Fieldwork in 1996 also involved the installation of 1) an
automated meteorological station, with a dipping bucket rain
gauge; 2) five peizometers for the measurement of groundwater-table fluctuations within the landslide; and 3) two
V-notched weirs to monitor runoff from the landslide and the
delivery of suspended sediments to Battle Creek. This
research is ongoing and, with the exception of the erosion pin
data, the field observations are preliminary and insufficient to
warrant reporting here. Therefore, the following results and
discussion are based on almost 2000 measurements of the
exposed lengths of the steel pins during May-October 1995
and May-July 1996.

D.J. Sauchyn and H.L. Nelson

GSC Bulletin 534

Between October 27, 1994 and July 3, 1996, net erosion


averaging 8.7 cm occurred at 84 pins (Table 2). Net deposition averaging 5.3 cm occurred at 15 pins and 2 pins showed
no measurable change. Fourteen pins were lost, either undermined by erosion or buried. Average net erosion of 6.4 cm
among all 101 pins represents 2290 m3 of sediment exported
from approximately 0.35 krn2 of earthflow deposits and secondary scarps. Maximum net erosion and deposition at single
pins were 58.4 and 22.2 cm, respectively, with 45 cm of erosion occurring at one pin as a result of a single storm in late
June 1995. During the same storm, most of the other pins
recorded little or no change, reflecting the importance of
gully erosion as the dominant mechanism of sediment production on the landslide. Net erosion has a typical, positively
skewed frequency distribution (Fig. 7).
The distribution of total erosion and deposition among the
20 observation days (Fig. 8) illustrates that the geomorphic
response to single rainstorms in May and June was comparable to, and even exceeded, the cumulative effects of precipitation and snowmelt runoff from late October to May, and
summer rainstorms in 1995. The pin observations were concentrated in May and June, when about 70% of the annual
precipitation occurs. The first measurements of 1996 (May 5)
revealed almost 2 m of total deposition and the maximum

single accumulation of 44.1 cm at one pin (Table 2). Then,


during the period May 5-15, with just 7 rnrn of rain, there was
more than 2 m of measured erosion, but negligible deposition.
Another 50 cm of erosion was measured just four days later.
During the rest of May and June, the effect of rainfall was
minimal. Similarly, the observations of May 5 and 10, 1995
indicate significant erosion but minimal deposition
(i.e. export of sediment from the monitored area and deposition beyond the landslide), following relatively light rainfalls
of 6.6 and 9.4 mm, respectively. Thus, the data in Figure 8
suggest a disproportionate effectiveness of frequent spring

Table 2. Summary of erosion and deposition at

101 erosion pins, October 27,1994to July 3,1996

Pins
Erosion
Deposition
No change
Net erosion

84
15
2
101

Mean
(cm)
8.7
5.3

Maximum1
(cm)
58.4
22.2

Maximum2
(cm)
45.4
44.1

6.4

'Maximum totals for one pin over the period of observation


'Maximum single observations

Figure 5. Morphological evidence of the activity and erosion of the Police Point landslide: a) A
lowfault scarp that formed at the base of the main scarp, which is about 10 m high (Fig. 3 ) in the
spring of 1992, when the uppermost slump block subsided about 30 cm. Photograph by
D.J. Sauchyn; GSC 1999-040B; b) Rill erosion of a proximal secondary scarp (Fig. 3).
Photograph by D.J. Sauchyn; GSC 1999-040C; c ) Debrisflow from the steep distal scarp (Fig. 3 )
and location of the erosion pins in tiers 9 & 10; erosional scarp on right is about 1.5 m high
(Fig. 6).Photograph by H.L. Nelson; GSC 1999-0400; d) Gully erosion of the earthjlow deposits
that constitute the lower part of the landslide (Fig. 3) and location of the erosion pins in tiers 1-8
(Fig. 6). Photograph by D.J. Sauclzyn; GSC 1999-040E.

D.J. Sauchyn and H.L. Nelson

0.0

10.0

20.0

30.0

40.0

50.0

60.0

Erosion (cm)
Figure 7. Positively skewed frequency distribution of net
erosion at 84 pins for the study period, October 27, 1994 to
July 3, 1996. Maximum net erosion at a single pin was
58.4 cm.

0 Deposition (cm)

Figure 6. Locations of the erosion pins on the earthflow


deposits (tiers 1-8) and a secondary scarp (tiers 9-10). Black
arrow points downslope (direction of flow). The aerial
photograph is from 1983. Photo courtesy of Alberta
Environmental Protection, Air Photo Services (photo no.
83-25, taken July 23, 1983).

May

June

Ib
Oct

0 Deposition (cm)

Figure 8. Total (cumulative) erosion and deposition


measured at the steel pins on: a) 11 days during the period
May 2 to October 15, 1995 and b) 9 days during the period
May 5 to July 3, 1996. Dates marked on x-axis are those on
which measurements were taken. Vertical bars give the
amount of rainfall (vnm)measured by Alberta Parks Service
staffnear Elkwater, Alberta, approximately 20 km west of the
landslide.

Date 1995

i3
Sept

1'5
May

1$

ia 3, i

2'7

Date 1996

June

3
July

GSC Bulletin 534

-6

70

60

50

40

30

a
T i
-Tier
-Tier

5-6
3-4

,z

e
W

20 10 0 2 May 5 May loMay l4May 24May 4 Jun 11Jun 2OJun 26 Jun 23Sep 150ct

Date 1995
T i
-Tier
-Tier
T i

CONCLUSIONS

r 7-8

r 7-8
5-6

r 1-2

3-4

5May 15May 19May 28 May 31 May 24 Jun

6 Jun 27 Jun

3 JuI

The Police Point landslide remains largely oonvegetated after


30 years. Uprooted tree seedlings suggest that revegetation
during periods of geomorphic inactivity has been thwarted by
debris flows, shallow landsliding, and gully erosion during
subsequent wetter years. Previous attempts to stabilize the
landslide by seeding brome grass were predictably futile
(cf. Clements et al., 1970; R.L. & L. Environmental Services
Ltd., 1994). Data presented here include 45 cm of erosion and
44 cm of deposition at individual pins during a single rainstorm. The variability among pins and storms suggests a
threshold response of surface sediments to rain and snowmelt
runoff, whereby much of the annual sediment production and
redistribution occurs during relatively few runoff events.
Ongoing monitoring of the erosion pins, as well as hydrological and meteorological parameters, should enable quantitative evaluation of these thresholds.
The Police Point landslide serves as a modern example of
the impact of landsliding on the major valleys of the Palliser
Triangle. Failure of plastic clays causes fracturing and displacement of overlying strata and exposes these sediments to
prolonged erosion and slow mass wasting. These processes
maintain the instability of landslide deposits for at least
decades and generate sediment that impacts on human activities and natural systems well beyond the valley sides.

Date 1996

Figure 9. Total erosion measured at pairs of adjacentpins on


the observation days indicated, illustrating the spatial and
temporal variation in sediment loss, during: a) I 1 days in the
spring of 1995, and b) 9 days in the spring of 1996. The pins in
tiers 1-8 and 9-10 are on the earthflow deposit and distal
scarp, respectively (Fig. 3, 6).

rain, and a highly variable geomorphic response to individual


storms determined by rainfall characteristics and antecedent
conditions. Higher magnitude rainfalls occurred from late
June to September, but precipitation was less frequent, water
tables were lower, and evaporation rates much higher; the
landslide therefore remained relatively dry and less responsive to hydroclimatic events.
The spatial variability of sediment redistribution (Fig. 9)
largely reflects episodic shallow mass wasting of the distal
scarp and headward growth of gullies into the earthflow
deposits (Fig. 3).The considerable variability between observation days and tiers of pins, as well as the skewed distribution of the erosion data (Fig. 7), further document the
dominance of gully erosion on the earthflow deposits (tiers
1-8; Fig. 6). It also precludes the significance of sheetwash
(as discussed by Clements et al. (1970)), which would produce a more uniform spatial distribution of sediment loss.
Tiers 9-10 are on a scarp (Fig. 6). where six of the
twenty-four pins have been lost in debris flows and shallow
slides. Thus, the data from these tiers underestimate the
degree of geomorphic activity. Measurements taken on
May 15,1996 record the greatest total erosion (summation of
all pins) measured in this study (Fig. 9).

ACKNOWLEDGMENTS
This research was supported by the Geological Survey of
Canada (Palliser Triangle Global Change Project Grant),
AlbertaFish and Wildlife, and Environment Canada, and was
conducted with the permission of the Alberta Parks Service.
Tamara Golemba, Scott Hunter, Jason Cosford, Linda
Shannon, Lowell Strauss, and Shannon Hall assisted with the
installation and measurement of the erosion pins. We thank
Don Lemmen and Jan Bednarski for thorough reviews of the
manuscript, and Dan Reesor of Graburn Gap Ranch for
accommodation and advice in the field.

REFERENCES
Christiansen,E.A.

and Sauer, E.K.


Age of the Frenchman Valley and associated drift south of the
Cypress Hills, Saskatchewan, Canada; Canadian Journal of Earth
Sciences, v. 25, p. 1703-1708.
Clements, S.H., Lees, V.W., and Atwood, L.
1970: Battle Creek habitat evaluation project: a preliminary investigation
of erosion control methods; Fish and Wildlife Division, Alberta
Department of Lands and Forests, Lethbridge, Alberta, 23 p.
de Lugt, J. and Campbell, I.A.
1992: Mass movements in the badlands of Dinosaur Provincial Park,
Alberta, Canada; in Functional Geomorphology, (ed.)
K-H. Schmidt and J. de Ploey; Catena, Supplement 23, p. 75-100.

1988:

Furnival, G.M.
1946:

Cypress Lake map area, Saskatchewan; Geological Survey of


Canada, Memoir 242, 161 p.
Janz, B. and Treffry, E.L.
1968: Southern Alberta's paralyzing snowstorms in April 1967;
Weatherwise, April, p. 70-75.94.

D.J. Sauchyn and H.L. Nelson


Kent, D.M.. and Simpson, F.
1973: Geological road log of the Cypress Hills-Milk River area, southeastern Alberta; in An Excursion Guide to the Geology of
Saskatchewan, (ed.) F. Simpson; Saskatchewan Geological
Society, Special Publication 1, p. 297-318.
Leckie, D.A. and Cheel, R.J.
1989: The Cypress Hills Formation (upper Eocene to Miocene): A
semi-arid braid plain deposit resulting from intrusive uplift;
Canadian Journal of Earth Sciences, v. 26, p. 1918-1931.
Lemmen, D.S., Vance, R.E., Campbell, I.A., David, P.P.,
Pennock, D.J., Sauchyn, D.J., and Wolfe, S.A.
1998: Geomorphic systems of the Palliser Triangle, southern Canadian
Prairies: description and response to changing climate; Geological
Survey of Canada, Bulletin 521, p. 48-54.
Lindoe, L.O.
1965: Ceramic clays of the Cypress Hills; in Cypress Hills Plateau,
Alberta and Saskatchewan, (ed.) R.L. Zell; Alberta Society of
Petroleum Geologists, Calgary, Alberta, 15th Annual Field
Conference Guidebook, Part 1, p. 210-225.
MacPherson, H.J. and Rannie, W.F.
1969: Geomorphic effects of the May 1967 flood in Graburn watershed,
Cypress Hills, Alberta, Canada; Journal of Hydrology, v. 9,
p. 307-321.

Mollard, J.D.
1977: Regional landslide types in Canada; in Landslides, (ed.)
D.R. Coates; Reviews in Engineering Geology, Volume 111,
Geological Society of America, Boulder, Colorado, p. 29-56.
R.L. & L. Environmental Services Ltd.
1994: Preliminary fisherics investigation of Battle Creek and Graburn
Creek, 1991-92; AlbertaFishand Wildlife, Report No. 306F, 32 p.
Sauchyn, D.J.
1993: Quaternary and Late Tertiary landscape evolution in the western
Cypress Hills; in Quaternary and Late Tertiary Landscapes of
Southwestern Saskatchewan and Adjacent Area, (ed.)
D.J. Sauchyn; Canadian Plains Research Center, Regina,
Saskatchewan, p. 4 6 5 8 .
Sauchyn, D.J. and Lemmen, D.S.
1996: Impacts of landsliding in the western Cypress Hills, Saskatchewan
and Alberta; in Current Research 1996-B; Geological Survey of
Canada, p. 7-14.
Thomson, S. and Morgenstern, N.R.
1977: Factors affecting the distribution of landslides along rivers in southern Alberta; Canadian Geotechnical Journal, v. 14, p. 508-523.
1978: Landslides in argillaceous rock, Prairie Provinces, Canada; in
Rockslides and Avalanches 2, (ed.) B. Voight; Elsevier Scientific
Publishing, New York, Developments in Geotechnical Engineering, 14B,p. 515-540.

Geomorphic surfaces and postglacial landscape


evolution of the Maple Creek basin, Saskatchewan

Vreeken, W.J., 1999: Geomorphic surfaces and postglacial landscape evolution of the Maple
Creek basin, Saskatchewan; in Holocene Climate and Environmental Change in the Palliser
Triangle: A Geoscientific Contextfor Evaluating the Impacts of Climate Change on the Southern
Canadian Prairies, (ed.) D.S. Lemmen and R.E. Vance; Geological Survey of Canada,
Bulletin 534, p. 267-294.

Abstract Maple Creek basin, a 75 km long closed drainage in southwestern Saskatchewan, contains
well differentiated, postglacial, fluvial and lacustrine geomorphic surfaces in addition to relict preglacial,
glacial, and glaciolacustrine surfaces. Deglaciation (ca. 18 000-13 000 BP) produced vast tracts of internally drained hummocky moraine and cross-valley glaciogenic barriers that gave rise to through-flowing
postglacial lakes. Postglacial Junction and Hanson lakes regulated flow to the terminus and functioned as
sediment traps and temporary base levels for stream erosion upvalley, until their demises at about 6000 and
1500 BP, respectively. The major geomorphic events controlling the postglacial evolution of the basin were
not climatically controlled. Rather, they reflect the delayed (up to 11 000 years) geomorphic response to
deglacial landscape instability. Direct response to Holocene climatic change is largely restricted to fluvial-runoff processes, such as alluvial-fan formation and hillslope erosion, that can remain unaffected by
base-level changes. Both show intervals of accelerated activity that correlate with the warm and dry
Hypsithermal climate interval (ca. 9000-4600 BP).

RCsumC : Le bassin du ruisseau Maple, bassin hydrographique ferm6 de 75 km de longueur dans le


sud-ouest de la Saskatchewan, prtsente des surfaces gtomorphologiques fluviales et lacustres
postglaciaires bien difftrencites, outre des surfaces prtglaciaires, glaciaires et glaciolacustres reliques. La
dtglaciation (environ 18 000-13 000 BP) a engendrt de vastes ttendues de moraines bosselCes I? drainage
endorCique ainsi que des barrages glaciogbnes transversaux de vallte qui ont donnt lieu ?I la formation de
lacs postglaciaires h tmissaires. Les lacs postglaciaires Junction et Hanson ont rCgulC 1'6coulement
jusqu'au front glaciaire et ont agi comme pibges ddimentaires et comme niveaux de base temporaires pour
1'Crosion fluviale en amont, jusqu'h Ieur disparition vers 6 000 et 1 500 BP, respectivement. Les principaux
Cvtnements gComorphologiques r6gissant 1'Cvolution postglaciaire du bassin n'ont pas 6tC contr81ts par le
climat. 11s traduisent plut8t la rtaction gtomorphologique retardte (jusqu'a 11 000 ans) a l'instabilitt des
paysages postglaciaires. La rtaction directe aux changements climatiques holocbnes se lirnite pour
I'essentiel a des processus d'tcoulement fluvial tels que la formation de c8nes alluviaux et 1'Crosion des
versants, qui ne sont pas ntcessairement touchts par des changements du niveau de base. Ces deux
phtnombnes tbmoignent d'intervalles d'activitt acctltrte qui sont en corrtlation avec l'intervalle
climatique chaud et sec de I'hypsithermal (environ 9 000-4 600 BP).

Department of Geography, Queen's University, Kingston, Ontario K7L 3N6

GSC Bulletin 534

INTRODUCTION
Understanding the geomorphic history of an area is prerequisite to understanding relationships between climate and
geomorphic response. Simple chronological correlation
between climate and geomorphic events does not indicate
causality, because the complexity and antecedent states of
geomorphic systems often result in delayed response to forcing factors. Closed drainages are ideal settings in which to
evaluate basin-scale landscape response because they are
excluded from 'upstream' geomorphic effects of base-level
changes external to them. Small closed basins reveal much
clearer relationships between climate change and process
response (e.g. Vreeken, 1996a) than larger ones, where such
linkages may be obscured by additional factors. This is relevant to predictions of the impacts of future climate change,
which will necessarily have decreasing confidence at
decreasing spatial scale, and particulwly with respect to
stream systems, which may show the least predictable
response (Lemmen et al., 1998).
Maple Creek basin, alarge closed drainage near the centre
of the Palliser Triangle, is ideally suited to evaluate the history of these associations. It features landforms denoting distinct geomorphic events and is located close to sites with well
controlled paleolimnological records of Holocene climate
(Fig. 1). This study examines the geomorphology, stratigraphy, sedimentology, and geochronology of the postglacial
basin within a longer term context of landscape evolution,
using a methodology based on mapping, landscape analysis,
and dating of geomorphic surfaces.

Study area
Maple Creek basin is a closed drainage, heading at 1227 m
a.s.1, in the Cypress Hills upland and terminating on the Sand
Hills-Bigstick Lake plain; it drops 518 m over its 75 krn
length (Fig. 1; Acton et al., 1960). Gap and McCoy creeks are
its main tributaries, with confluence in the Triple Junction

area immediately north of the town of Maple Creek, while


Bigstick and Bitter lakes serve as terminal distributaries
(F;~. 2). Underlying bedrock is subhorizontal, ranging from
marine shale of the Upper Cretaceous Bearpaw Formation
(subcropping below approx. 1190 m), through sandstone,
siltstone, and lignite, to fluvial gravel and sand of the Upper
Tertiary Cypress Hills Formation, which outcrop in the basin
headwaters (Klassen, 1991).
The Late Wisconsinan Laurentide Ice Sheet covered virtually the entire basin and extended across saddles on the continental divide, but the headwater plateaus remained
nonglaciated (Fig. 3). Deglaciation began ca. 18 000 BP, with
meltwater drainage initially occurring southward across the
continental divide (Christiansen, 1979;Dyke and Prest, 1987;
Klassen, 1994). Subsequent retreat featured large areas of ice
stagnation, with meltwater accumulating in progressively
lower proglacial lakes that drained along the north flank of the
upland (Klassen, 1994). With retreat from the Fox Valley
moraine, north of Bitter and Bigstick lakes, shortly after
13 000 BP, the active ice margin lay north of the basin terrninus (David, 1964; Christiansen 1979; Dyke and Prest, 1987).
Bigstick and Bitter lakes (Fig. 2,4) are themodern successors of the deglacial drainage systems, while the hummocky
surfaces remain mostly internally drained. As a result, the
perimeter of the upper basin greatly exaggerates the actual
flow- and sediment-contributing area. Because the lower
basin has virtually no tributaries, its perimeter follows the
valley wall (Fig. 2).
The basin has a steep climatic gradient from subhurnid to
semiarid. Mean annual precipitation ranges from 450 rnrn in
the Cypress Hills to 300 rnm on the plains to the north. Similarly, from south to north, the frost-free period ranges from
less than 60 to about 110 days, mean annual potential evaporation from 5 10 to more than 560 mm, and mean annual moisture deficit from less than 100 to greater than 200 mm
(Canada Centre for Mapping, 1990). The prevailing wind
direction is from the west. Vegetation ranges from
spruce-lodgepole pine forest at the headwaters to low-grass

SAND HILLS IGSTICK LAKE

Medicine Hat

Aspen parkland
Bunchgrass steppe
Northern
mixed-grass prairie

0
Ham site
49"

km

110"

Figure 1. Setting of Maple Creek basin: A) native vegetation of the southern Canadian Interior; B) location
relative to continental divide and other study sites mentioned in text (indicated by black circles).

108"

W.J. Vreeken

Figure 2. Topography of Maple Creek basin. Dashed line delineates subaerial catchment of
upper basin. Solid circles denote study sites (see Fig. 3 for details). Elevations in metres
above sea level. Maple leaf indicates Trans-Canada Highway.

Bulletin 534

110" OO'W
Cypress Plain
Upper Gap
glacial driftplain
Downie glaciolacustrine surface
Neitz glaciofluvial surface

J' !J!p

@ Unnamed preglacial
erosion surfaces
Lower Ga
glacial driiPlain
Mapje Creek.

EBI

Dissected bedrock
surfaces

. .r.

Unnamed glacioIfluvial
n plains
e

Units 13-19
of Table 1

109" 13'W
Blacker Lake
glacial driftplain
Unnamed morainal
surfaces complex
Unnamed glacio-

flfh.Aib,";mmocky

W.J. Vreeken

Figure 3. Geomorphic su$aces, unnamed terrain complexes,


major meltwater channels, and study sites in the Maple Creek
area. Delineations modified after Saskatchewan Research
Council (1987) and Klassen (1991). Dashed line outlines
subaerial catchment of upper basin. Circled numbers denote
study sites.

,Cypress Plain
UPPER MAPLE CREEK

' /I

Upper Gap surface

~~~i~ channel

channel
Mc~Ougald-~<]-

Lake Downie plan

Lake Maple Creek pialn

GAP CREEK

GI

iI
o

600

complexes in Maple Creek basin1.

1 Code / Unit name / Characterization

Postglacial units:
19

Bigstick

Modern lake plain

18
17

Bull

Modern alluvial plain and channel

Hanson

Lake plain

16
15
14

Weir

Alluvial plain

Gapiet

Alluvial plain

Lawrence

Subaerial alluvial fans

13

Junction

Fan delta (f), lake plain (I), and lake outlets (0)

Glacial and deglacial units:

7 Bedrock slopes

\+,

Table 1. Geomorphic surfaces and unnamed terrain

~igstick/~ake
plain
LOWER MAPLE
CREEK
I
I
I
I

12

Neitz

Glaciofluvial surface

11

Unnamed

Glaciofluvial plains (p) and hummocky (h)


surface complexes

10

Maple Creek Glaciolacustrinesurface

Downie

Glaciolacustrinesurface

Unnamed

Morainal surfaces complex

Unnamed

Dissected buried preglacial erosion surfaces


complex

Lower Gap

Glacial driftplain

Upper Gap

Glacial driflpiain

Blacker Lake Glacial driftplain

Unnamed

Dissected bedrock surfaces complex

Preglacial units:

Distance from confluence (km)

erosion surfaces complex

Figure 4. Longitudinal stream profiles for Maple and Gap


creeks, in relation to geomorphic features.
grasslands on the plains (Clayton et al., 1977). Well drained
soils range from Black and Dark Brown to Brown
Chernozemic, with Solonetzic, Regosolic, and Gleysolic
inclusions (Clayton et al., 1977; Saskatchewan Institute of
Pedology, 1990, 1991). Most land is presently used for grazing, dryland, and irrigation farming, enabled by four
water-storage reservoirs constructed since 1938: Downie
Lake (12.2 x lo6 m3, 1938), Junction Reservoir (10.9 x
lo6 m3, 1939), McDougald Reservoir (0.9 x lo6 m3, 1940),
and Harris Reservoir (6.0 x lo6 m3, 1956). As a result,
Bigstick Lake receives little streamflow today and Bitter
Lake receives flow only from rare floods.

METHODOLOGY
The foundation of this study is the identification and mapping
of geomorphic surfaces, which reflect the succession of
preglacial, glacial, and postglacial geomorphic systems that
operated in the study area (Table 1). Formally defined by
Ruhe (1975) and Vreeken (1984), a geomorphic surface or
complex of surfaces is 1) a part of the land that is specifically
defined in space and time and may include many landforms;
2) a mappable feature whose geographic distribution is

' Codes generally identify map units shown in Figures3, 6, 7, 8, and 9.

In Figure 3, units 13 to 19, inclusive, are collectively coded as unit 20.

portrayed on maps or airphotos and whose geometric


dimensions are specified and analyzed; 3) associated with
other geomorphic surfaces which are defined in order to place
it in a space and time sequence; 4) associated with rock or
sediment below it, or on it, which must be specified; 5) dated
by relative or absolute means, which must be specified; and 6)
labelled with a geographic name. A geomorphic surface may
be erosional or depositional, or both.
Geomorphic analysis was undertaken using elevations
taken from 1:50 000 scale NTS topographic maps with
25-foot contour intervals, and from an unpublished 1:9150
map with selected contour lines for the Triple Junction area
(Fig. 2; Prairie Farm Rehabilitation Administration (unpublished data, 1938)). Field descriptions of sediments utilized a
lithofacies notation (Appendix A) adapted from Mia11 (1978,
1992) and Eyles et al. (1983). Soil descriptions followed
standard pedological terminology (Canadian Agricultural
Services Coordinating Committee, 1983). Relative ages of
surfaces and sediments were inferred from spatial relationships (Ruhe, 1975; Vreeken, 1984), while absolute ages were
determined by radiocarbon dates (Table 2) and tephra beds.
The identities of a Glacier Peak tephra bed (1 1 200 BP; Foit
et al. (1993)) at site 7 (Fig. 3), and of Mazama tephra beds

GSC Bulletin 534

(6800 BP; Bacon (1983)) at sites 2,3,6, 10, and 15 (Fig. 3)


were established from electron microprobe determination of
macroelemental compositions. Tephra beds from postglacial
sequences at other sites are taken to represent the Mazama
tephra fall.

RESULTS
Geomorphic surfaces or surface complexes defined for the
Maple Creek basin are listed in Table 1 and portrayed in Figures 3,6,7, 8, and 9. Spatial, sedimentary, and geochronological attributes and associations are presented below.
Sedimentological descriptions and interpretations for 19 sites
(Fig. 3) are provided in Appendix A.

Preglacial surfaces
Cypress Plain (code 1, Fig. 3) and undifferentiated
erosional surfaces (code 2, Fig. 3)
Remnants of Cypress Plain (Alden, 1932), the highest
preglacial surface between the Rocky Mountains of Alberta
and the Torngat Mountains of Labrador, are confined to three
plateaus along the basin headwaters and delimited by large
south-directed meltwater channels (Klassen, 1991). Four
south-sloping erosion surfaces, angularly juxtaposed,
descend from the divide, the highest of them bevelling the
Cypress Plain (Vreeken et al., 1992). Although dissected and
partly buried, their moi-phology remains distinct, occupying
interfluves between integrated tributaries. Contours of comparable north-sloping surfaces (code 7, Fig. 3) and a buried
axial valley beneath lower Gap Creek (Klassen, 1991) reflect
the preglacial origin of the Maple Creek basin (see also
Klassen, 1992).

Table 2. Radiocarbon dates from Maple Creek basin.

Lab no.

Date
(years BP)

+ 340

Material

Landscape
setting
(elevation)

Location (site no.);


latitude, lonaitude

Shells
(Stagnicola)

Downie surface
(833 m)

Fleming Creek (site 1);


4g049'20 N, 109O29'00 W

Downie surface
l(833 m)
Downie surface
(823 m)

Fleming Creek (site 1);


4g049'20 N, 10929'00 W

GSC-4675

14 000

TO-694

13 120 f 80

Shells
(Stagnicola)

TO-5279

11 780 f 70

Shells
(Siagnicola)

TO-5280

10 510 f 70

Charred wood

* 80

TO-4413

9520

TO-201

8590 f 100 Shells

TO-4410

7480 f 80

Low-organic
sediment

GSC-4422

7270f80

Wood
(Salix)

GSC-6068

7250

TO-4411

4710 k 70

Low-organic
sediment

TO-4411-R

6240 f 70

Low-organic
sediment

TO-100

3600f 80

Bone

GSC-4027

2940 f 60

Shells
(Sphaerium)

TO-4415

2280f 50

Bone

TO-4416

1520 k 50

Bone (bison)

TO-4414

110f50

+ 100

Low-organic
sediment

Wood

Bone

GSC-4421
(Salix)

Comments
Depth of 2 m in glaciolacustine sediments

Friday 106 (site 7);


49"51f00 N,1093335 W

Depth of 9.6 m in postglacial depression fill;


Glacier Peak tephra 0.7 m lower

Downie surface
l(823 m)
Bull surface
(715 m)

Friday 103 (site 7);


49"51'00 N,10g033'55 W

Depth of 9.3 m in postglacial depression;


TO-5279 is 0.3 m lower

Feil's crossing (site 18B);


50'1 1'00 N,10g027'38 W

Depth of 5.9 m below alluvial surface

Gap surface
I(1112m)

Black (site 2);


4g037'1 0 N,109'42'30 W

Depth of 3.2 m in 7 m thick postglacial fill


on till

Bigstick surface
(715 m)

Bigstick 16 (site 19E);


5014'34 N,109"24'00 W

Depth of 2.1 m below lacustrine surface

Lawrence surface
(803
Lawrence surface
(803 m)
Bigstick surface
(715 m)

Gap Creek 2 (site 6);


49'51'05 N,109"35'25 W

Stream sediments 9 m below fan surface

Gap Creek (site 6);


49'51'05 N,109'35'25 W

Stream sediments 9 m below fan surface;


compare with GSC-4422

Bigstick 16 (site 19E);


50'1 4'34 N,109"24'00 W

Depth of 5.2 m below lacustrine surface

Bigstick 16 (site 19E);


50'14'34 N.1 0924'00 W

Repeat of TO-4411

Gap Creek, The Weir (site 8);


4g051'33 N,109"35'00 W

Depth of 2 m in sandy alluvium

Gap Creek 1 (site 8);


49'51'33 N,109"35'00 W

Depth of 2 m in sandy alluvium;


compare with TO-100

Junction Reservoir 1B (near


site 10);
4g055'55 N,10930'30 W

Depth of 0.3 m in stream alluvium that


postdates Junction Lake plain

1
I

Bigstick surface
1 ~ 1 m)
5
Weir surface
(785 m)

I
1

Weir surface
(785 m)
Junction surface
(760 m)
Weir surface
(798 m)
Junction surface
(760 m)

Gap Creek 5 (near site 8);


49'50'57 N, 109"34'00 W

Depth of 0.5 m in loam beneath alluvial


surface

Junction Reservoir 2
4g056'00 N,10930'30 W

Depth of 0.7 m below lacustrine surface

Bull surface
(792 m)

Gap Creek 3
49'51'05 N,109"35'25 W

Hearth, 0.3 m below top of sandy loam


alluvium

W.J. Vreeken

Cypress Plain records a braidplain at the top of the


Cypress Hills Formation (Leckie and Cheel, 1989). The
younger erosion surfaces are overlain by thin gravel units
identical to those beneath Cypress Plain (Vreeken, 1986;
Vreeken et al., 1989). These preglacial surfaces are locally
buried beneath Davis Creek silt, sandy loam, silt loam, and
clay loam deposits that include seven tephra beds (Vreeken
et al., 1989, 1992). Strongly developed reddish brown
paleosols formed from these loam deposits, and elsewhere
from gravel beds, have strong overprints from cryoturbation.
Planar erosional surfaces cut the cryoturbates and are overlain by one or more of the following: noncryoturbated gravel
with Canadian Shield rock types, diarnict interbedded with
silt-clay rhythrnites, and loess that includes a tephra bed,
likely Mazama. Comparable overburdens have not been
found on north-sloping erosion surfaces from this study area,
but occur near Elkwater, Alberta (Vreeken, 1986).
Relative ages of the preglacial surfaces are indicated by
their geometric relationships and relative height, with
Cypress Plain being the oldest and the four progressively
lower erosion surfaces being successively younger. The maximum age of Cypress Plain is Middle Miocene, based on the
youngest fossils from its substrate (Storer, 1975). The age of
the Davis Creek silt is inferred to be Middle to Late Miocene,
based upon three paleomagnetic reversals within the silt and
an 8.3 Ma fission-track age of its uppermost tephra bed
(Vreeken and Westgate, 1992;Barendregt et al., 1997).These
data indicate that all the preglacial surfaces are of at least Late
Miocene age. The paleosols are obviously younger, but predate the cryogenic interval that modified them. The most
recent cryogenic interval in the western Cypress Hills lasted
from ca. 22 700 to maximally 10 300 BP (cf. Young et al.,
1994; W.J. Vreeken, S.A. Wolfe, and D.J. Huntley, unpub.
data, 1999).

Figure 5. Geomorphic ~ ~ $ a c eon


s the continental divide,
viewed from site 2 toward Cypress Hills centre block.
Cypress Plain and the Upper Gap and Lower Gap su$aces
are separated by dissected bedrock su$aces (DSB). Photograph by W.J. Vreeken. GSC 1999-043A

Downie surface (code 9, Fig. 3,9)

Glacial surfaces
Blacker Lake surface (code 4, Fig. 3) and Upper and
Lower Gap surfaces (codes 5 and 6, Fig. 3)
Glaciogenic surfaces in Maple Creek basin include the
Blacker Lake and Gap lobate driftplains on the continental
divide, as well as an undifferentiated morainal complex (code
8, Fig. 3) on the adjacent plains. These are areas of predominantly low- to high-relief hummocky topography and are
largely internally drained (Fig. 5).The Blacker Lake and Gap
drift lobes are flanked by major and minor meltwater channels that cross the continental divide. One of the smaller channels, presently occupied by Gap Creek and herein named Gap
spillway, separates the Upper and Lower Gap driftplains
(Fig. 3). Sediments beneath all of these glaciogenic surfaces
consist of silt loam and loam diamict (till; Klassen (1991,
1992)), and were not examined as part of this study. The
Blacker Lake and Upper Gap driftplains, being the highest,
were the first to become ice free and subaerially exposed. The
Lower Gap driftplain became exposed after surface drainage,
initially across the divide (Fig. 3), had been deflected along
the northern slope. Exposure of lower glaciogenic surfaces
was delayed by melting of stagnant ice and drainage of any
deglacial lake that extended across them.

The plain of glacial Lake Downie forms a generally


north-sloping, high-relief hummocky surface between 907
and 81.5 m a.s.l., extending southwest to Downie meltwater
channel (incised from approximately 960 to 938 m a.s.1. at the
limit of the drainage basin) and southeast to McDougald
channel (max. 861 m a.s.1.; Fig. 3). Along its northern limit,
the land surface drops sharply into the large topographic
basin of the Triple Junction area, which includes the largely
buried glacial Lake Maple Creek plain (see 'Maple Creek
surface').
Glacial Lake Downie sediments attain an exposed thickness of approximately 50 m along Gap Creek valley, and consist of planar crossbedded silt and sand, variably interbedded
with diamict and with thin clay beds near the top of sequences
(sites 1,3-9, Fig. 3; Appendix A). Beds in the upper sequence
are commonly inclined, folded, and normal faulted, and are
interpreted to reflect collapse upon melting of buried ice. The
uppermost diamict bed at site 1 (831 m a.s.1.) contains shells
of Stagnicola dated at 14 000 and 13 100 BP (GSC-4675 and
TO-694, Table 2). As these are shallow-water gastropods, it is
probable that subaerial conditions prevailed on a large part of
the Downie surface (907-815 m) by ca. 13 000 BP. Thepresence of Glacier Peak tephra in the uppermost lake sediments

GSC Bulletin 534

(813 m a.s.1.) at site 7 suggests that the lake plain drained


shortly after the tephra fall. The assumed age for distal Glacier
Peak tephras is 11 200 BP (Foit et al., 1993),but the optical date
for one such bed, near Elkwater Lake (Fig. I), is 12 000 BP
(radiocarbon age equivalent; W.J. Vreeken, S.A. Wolfe, and
D.J. Huntley (unpub. data, 1999)). The accumulation and preservation of this tephra bed within ripple-laminated fine sand
suggests that vigorous flow had ceased, and that the demise of
glacial Lake Downie and exposure of its entire lakebed were
imminent by ca. 12 000 BP. This time frame is supported by a date
of 11 800 BP (TO-5279, Table 2) on Stagnicola shells from a
closed basin fill on the Downie surface (site 7, Fig. 3; Appendix
A). A minimum date on subaerial exposure of this site, at 10
500 BP, is from charred wood from the first soil formed from
oxidized loess (TO-5280, Table 2).
Maple Creek surface (code 10, Fig. 3,6)
The plain of glacial Lake Maple Creek lies immediately north
of the Downie surface and is partly buried beneath the
postglacial Junction Lake plain (Fig. 6; see 'Junction surface
complex'). The paleoshore of glacial Lake Maple Creek lies
about 765 m a.s.l., and is marked by cobble-strewn surfaces that
slope away from notches in adjacent hillsides, near site 12
(Appendix A) and along the east shore of Junction Reservoir.
The northern limit of this plain is a ridge crossed by two spillways (Eagle and Lightfoot outlets), 0.8-1.0 km wide on the
saddles, the western one occupied by Maple Creek (Fig. 7).
Terrain north of this ridge, along lower Maple Creek, is lower
than the lake plain and spillways, and contains no connecting

meltwater channels (Fig. 2). Sediments associated with glacial Lake Maple Creek are exposed along the Junction
Reservoir bluffs (sites 12-14, Fig. 3; Appendix A). No
absolute dates are available for the sediments or associated
lake plain.
Neitz surface (code 12, Fig. 3,8) and Bitter Lake
meltwater channel
The low-relief Neitz surface (716 m a.s.1.) is underlain by
coarse sand and gravel deposits, interpreted as glaciofluvial,
that have no glaciolacustrine or postglacial lacustrine or fluvial overburden on them. It includes large southwest- and
south-directed flutes, and crescentic and comma-shaped
troughs resembling forms eroded by high-energy water
flow (Allen, 1984; Kor et al., 1991). Their trends suggest
that they originated from initially westward flow that turned
south, directed towards glacial Lake Maple Creek. Bitter
Lake meltwater channel extends west from near Tenaille
Lake, running perpendicular to and cut slightly below the
erosional landform assemblage. The channel continues
across the locations of Bitter and Many Island lakes (Fig. 2,
3), where its bottomland has prominent west-directed crescentic troughs. The Neitz surface is separated from the adjacent and higher Fox Valley morainal surface by along, 15 m
high, arcuate scarp and from the lower postglacial Bigstick
Lake plain by scalloped scarps (Fig. 8). Subaerial exposure
occurred after the last westward meltwater discharge, by
which time the Fox Valley moraine and all terrain to the
south of it were likely ice free.

Figure 6.
Geomorplzic surfaces and study sites in the
Triple Junction area. Circled numbers denote
study sites. Double line with cross-hatches
trending northwest through centre of area is
the railroad. Map derived from airphoto
(CSMA No. 801 81 -01 - L l l - 5 8 ) provided
courtesy of Saskatchewan Central Surveys
and Mapping Agency.

Unnamed morainal
surfaces complex
Junction Lake
pla~n

Maple Creek.

Unnamed glacio-

Gaplet alluvial
plain

Weir alluvial
plain

~ p ~ ~ ~ , " c u s tfluvial
r ' nhummocky
e~

Junction
fan delta
Bull modern
(181alluvial plain

W.J. Vreeken

Postglacial surfaces
Junction surface complex (code 13, Fig. 6,7, and 9)
hi^ complex, confined to the basin of glacial ~~k~ ~~~l~
Creek, includes a fan delta, a lake plain, and two outlets. The
fan-delta surface (785-759 m a.s.1.) issues from Gap Creek
valley and merges with the lake plain (Fig. 6), which extends

up McCoy and upper Maple creeks. Junction Lake was


dammed to the north by the same ridge that confined glacial
Lake Maple Creek, and its outlets occupy the same spillways
that controlled its predecessor (Fig. 7). The Lightfoot spillWay (765 m a.s.1.) is incised about 6 m by an outlet connected
to the Maple Creek valley. North of Junction Reservoir, incision of the Eagle spillway (also 765 m a.s.1.) is recorded by
successive erosional surfaces below the Junction Lake plain
(site 15, Fig. 7; Appendix A). Southeast of the reservoir dam,
a relict channel (758 m a.s.1.) is directed towards the Eagle
outlet (site 14, Fig. 7; Appendix A).
High stands of Junction Lake at 759-760 m a.s.1. are inferred from basal elevations of subaerial deposits overlying
glaciofluvial and glaciolacustrine deposits at the north end of
Junction Reservoir and near the spillways (sites 12, 13, 16,
Fig. 6, 7; Appendix A). Five major sedimentation cycles
within Junction Lake are reflected in 3 4 m high cliffs along
the shore of Junction Reservoir (sites 10 and 11, Fig. 6;
Appendix A). The environment during the last four of these
cycles likely was perennially lacustrine between these sites,

Unnamed glaciofluvial~ummocky

b
,J[%~$%%ine
Unnamed glacio-

bdJunction
Lake outlets
Lake

onit;;;

2:;:

Lake

alluvial

Bull modern
alluvial plain

Figure 7. Geomorphic surfaces and study sites along the middle Maple Creek reach. Circled numbers denote study sites.
Map derived from airphotos (CSMA No. A27727-66(S) and
12(N))provided courtesy of Saskatchewan Central Surveys
and Mapping Agency.

1181

Unnamed morainal
Unnamed, glaciofluvial lalns
sufiaces cornpiex 'Ip
Bull modern
aigstic! modern
alluv~alpla~n
lake plain

AT-

Neitz glaciofluvial surface


erosional
scarp

Figure 8. Geomorphic sugaces and study sites in the Bigstick


Lake area. Circled numbers denote study sites. Map derived
from airphotos (CSMA No. A27728-35(S) and 88(N))provided courtesy of Saskatchewan Central Surveys and Mapping Agency.

GSC Bulletin 534

following an emersive phase evidenced by a buried surface


that is marked by desiccation cracks filled with slickensided
clay at site 11. Sand influxes from the fan delta spread over
the lake floor as low-density turbidity currents, probably
reflecting high stream-discharge events. Sand influxes from
Maple Creek were confined to transient minor channels. An
erosive phase on the lake plain, following deposition of the
uppermost clay bed, was followed by accumulation of
clifftop loam, interpreted as loess.
Sedimentation in postglacial Junction Lake began
directly after meltwater flow from the north ceased entering
glacial Lake Maple Creek from the Bigstick Lake area.

Mazama tephra within lake-basin sediments (site 10, Fig. 6;


Appendix A) indicates that the lake plain and fan-delta surface postdate 6800 BP. The fan-delta surface is coeval with
the northernmost Lawrence fan that merges with it (see 'Lawrence surface complex'). Within the Lightfoot spillway
(759 m a.s.l.), reworked tephra grains at 0.6 m depth at site 16
(Appendix A) demonstrate that the outlet still transmitted
flow after the tephra fall. Lake level had to be about 759 m
a.s.1. for flow to reach the lower Maple Creek basin, so the
emersion surface at site 11 (approx. 755 m a.s.1.) indicates
that the lake basin was closed, with a water level more than
4 m below outlet level sometime before 6800 BP. A subsequent rise in lake level and renewed sedimentation resulted in
aggradation to within about 2 m of maximum water storage
level. By that time, downcutting of the Eagle outlet produced
the channel relict at site 14 (Appendix A), causing abandonment of the Lightfoot outlet and subsequent drainage of
Junction Lake.

Lawrence surface complex (code 14, Fig. 9)


Relict alluvial fans and fan aprons of the Lawrence alluvial
fan complex are the highest fluvial surfaces above the modern
floodplain (Bull surface). This complex descends about 50 m
from the upland of the Downie Lake plain to the lower Gap
Creek valley floor, commonly extending along and below
slump scarps along the valley walls. Channels leading into
fan apices extend only short distances onto the upland. The
northernmost fan grades into the southwestern part of the
Junction fan delta (code 13f, Fig. 9), but all other fan
toeslopes are truncated and stand above Weir and Bull surfaces, which occupy the remainder of the valley bottom
(Fig. 9, 10).

Unnamed morainal
surfaces complex

Downie glaciolacustrine surface

p
J alluvial
Lawrence
fans

$:%
:;

Bull modern
alluvial plain

alluvial
r(h

slump scarp

Figure 9. Geomorphic sutj5aces and study sites along lower


Gap Creek. Circled numbers denote study sites. Map derived
from airphoto (CMSA No. 80181-01-Ll0-178) provided
courtesy of Saskatchewan Central Surveys and Mapping
Agency.

Figure 10. Geomorphic surfaces along lower Gap Creek


near study site 6, looking southwest from site 8. Lawrence
alluvial fans descend the Downie Lake plain upland, their
truncated toeslopes standing above the Bull alluvial s u ~ a c e
and the terrace-forming Weir alluvial surface. Photograph
by W.J. Vreeken. GSC 1999-043B

W.J. Vreeken

Cutbank exposures at sites 4,5, and 6 (Fig. 9; Appendix A)


reveal fan deposits, 0.6-7.3 m thick, overlying loess (sites 4
and 5 ) or alluvial deposits (site 6). A tephra bed, probably
Mazarna, occurs within the loess at two sites and an identified
Mazama bed occurs within fan deposits at 5.3 m depth at site
6, indicating that the fan surfaces at all three sites are younger
than 6800 BP. Wood from buried stream sediments at site 6
was dated at 7300 BP (GSC-4422 and 6068, Table 2), indicating that fan aggradation prior to deposition of Mazama ash
averaged 0.4 m.100 a-l. Assuming uniform deposition rates
for the entire unit, this fan would have stabilized ca. 5500 BP
and formed in less than 1800 years.
Gaplet surface (code 15, Fig. 6)
The Gaplet alluvial surface issues from Gap Creek valley,
where it is slightly higher than the adjacent Weir surface. It
traverses the 1-2 m higher Junction fan delta and lake plain.
Its faint scrollbars and the outline of its margins indicate that
it resulted from meandering flow. As it is not traversed by the
modern floodplain (Bull surface), no exposures were available for study. The Gaplet surface postdates the Junction and
Lawrence surface complexes, but predates the Weir surface
(see below). I interpret this surface as an early product of the
Weir stream cycle, marking an abandoned course or
distributary of the trunk stream.
Weir surface (code 16, Fig. 6,7,9) and Hanson surface
(code 17, Fig. 7)
The terrace-forming Weir alluvial surface appears approximately 3 km downvalley from the confluence of Gap and
Downie creeks (861 m a.s.1.; Fig. 2) and grades into Hanson
Lake plain (732 m a.s.1.;Fig. 7). It has minor representation at
the low end of the upper Maple Creek reach but none along
McCoy Creek (Fig. 6). Weir terraces are 2-4 m high along the
modern floodplain (Bull surface; Fig. 10, 11) and lie 2 4 m
below truncated Lawrence fans and the Junction fan delta and

Figure 11. Geomorphic su$aces at debouchure of lower Gap


Creek valley onto Triple Junction lowland, looking
north-northeast from site 6. Upland near site 8 represents
Downie Lake plain. In the valley, Weir alluvial terraces stand
above BullJZoodplainand channel but they are set below the
Junctionfan delta and lake plain in thefar distance. Photograph
by W.J. Vreeken. GSC 1999-043C

lake plain. Its outline suggests that the surface is the product
of meandering flow. Hanson Lake plain is connected to several smaller lake plains to the east, with paleoshorelines
marked by cobble-and-boulder lags extending below hillside
notches. Its main outlet was the Ford outlet, now occupied by
Maple Creek (Fig. 7). A poorly recognizable older outlet was
connected to Tenaille Lake (Fig. 2,7). Hanson surface forms
a 2 m high terrace along the modern floodplain (Bull surface).
Sediments of the Weir terrace, overlying diarnict at site 8
(Appendix A), record a succession from channel gravel,
through lateral accretion deposits and overbank clay, to loess
(Fig. 9). A manmade channel through Hanson Lake plain
shows that a glaciofluvial substrate is overlain by postglacial
lacustrine clays with a fluvial sandy interbed (731 m a.s.1.;
site 17B, Fig. 7; Appendix A). Islands and shores associated
with the lake plain are underlain by glaciofluvial sediments,
and the absence of lacustrine deposits on them, at site 17A
(approx. 732 m a.s.1.; Appendix A), indicates that Hanson
Lake never stood more than about 732 m a.s.1.
Weir surface stands below, and hence postdates, the
coeval Lawrence and Junction surface complexes. Because it
was characterized by a meander belt up to 1 krn wide (Fig. 9),
a single succession of its lateral accretion deposits spans a
wide age range. Bone fragments and shells from lateral-accretion sand deposits at site 8 (Fig. 9) were dated at
3600 and 2900 BP respectively (TO-100 and GSC-4421,
Table 2; Appendix A), but may significantly postdate the
beginning of the cycle. Bone from overbank clay near site 8
was dated at 1500 BP (TO-4416, Table 2), while sediments
from a minor tributary, incised below Junction Lake plain
(site 10, Fig. 6; Appendix A) were dated by a bone at 2300 BP
(TO-4415, Table 2). The conclusion is that the Weir stream
cycle began between 6800 and 3600 BP, and that it ended ca.
1500 BP with the demise of Hanson Lake. Accordingly, the
Weir and Hanson surfaces date to about 1500 BP.

Bull surface (code 18, Fig. 6,7,8,9)


Bull surface is the expression of ongoing channel and
floodplain processes. Below the junction of Gap Creek and
the Downie Creek meltwater channel, Bull surface is accompanied by relict fluvial surfaces described previously, and
irregularities along its longitudinal profile are associated with
the Junction Lake and Hanson Lake plains (Fig. 4). In the
lowermost reach, it grades into the Bigstick Lake and Bitter
Lake plains. Near the transition from Bull surface to Bigstick
surface, at Feil's crossing (715 m a.s.1.; site 18, Fig. 8;
Appendix A), sediments suggest that initial lacustrine conditions, extending upvalley from Bigstick Lake, were followed
by recurrent fluvial influence and occasional shoaling, possibly emersion. Numerous thick clay beds within the upper 4 m
at site 18B record recurrent upvalley extension of Bigstick
Lake, up to 6 km from Feil's crossing but constrained by a
maximum elevation of 716 m a.s.1.
Because incipient formation of the Bull surface postdates
the youngest adjacent geomorphic surface, it dates minimally
to completion of the Weir cycle (ca. 1500BP) and maximally
to deglaciation, both in the upper Gap Creek reach and
between the Hanson Lake and Bigstick Lake plains. Charred

GSC Bulletin 534

wood from overbank clays at site 6 (Fig. 3; Appendix A),


adjacent to a Weir terrace, was dated at 40 f 50 BP
(GSC-4421, Table 2), thus confirming the modern age.
Bigstick surface (code 19, Fig. 8)
Both the Bitter Lake and Bigstick Lake plains merge with the
modern floodplain (Bull surface). The Bitter Lake plain was
not examined in this study. The limit of the Bigstick Lake
plain was taken at a neck of the Maple Creek floodplain, at
Feil's crossing (Fig. 8). The Bigstick inlets area, between this
point and the lower of two dams, includes shallow dry channels between domed flats and lies below remnants of the
Neitz surface. Smooth clay predominates below 2 m depth in
the western inlet channel (site 19E, Appendix A), confirming
that initial lacustrine conditions preceded recurrent channel
influence, as at site 18B.
Postglacial sedimentation in Bitter and Bigstick lakes
began simultaneously, after the last discharge of meltwater
from Bigstick basin. Early postglacial Bigstick Lake
extended significantly up the present Maple Creek valley.
Organic material, underlain by more than 2.5 m of lacustrine
clay from site 18B (Appendix A) and overlain by 1.2 m of
clay, was dated at 9500 BP (TO-4413, Table 2). The top of a
similar lacustrine clay sequence at site 19E (Appendix A) was
dated at 7500 BP (TO-4410, Table 2). This suggests that
Bigstick basin received ample streamflow until ca. 7500 BP
and that, upstream, Junction Lake likely did not experience
prolonged low water levels. The mean sedimentation rate at
site 18B would have been 0.06 m. 100 a-l, between 9500 and
7500 BP, although initial sedimentation was probably faster
than the calculated mean rate.
Other sites
Important records of postglacial environmental change are
also provided by sites that do not occur on postglacial
geomorphic surfaces. A closed-basin fill on Upper Gap surface (site 2, Fig. 3; Appendix A) includes 5.5 m of slope-wash
sediments with shells and ostracode assemblages indicating
that perennial pond conditions ended after 8600 BP, and that
parkland on the continental divide changed to prairie between
8600 and ca. 7000 BP (L.D. Delorme, unpub. data, National
Water Resources Research Institute, Service Report
88-0-219,1988; Vreeken, 1993a, 1994). Shallow closed basins on Downie surface, around Harris Reservoir (Fig. 2), are
filled with sandy loam, interpreted as loess, with buried soils;
one of these buried soils contains tephra grains, likely
Mazama. A deep enclosed basin at site 7 (Appendix A) contains the thickest postglacial loess sequence in the basin (9.2
m), and includes 40 buried soils and a Mazama tephra bed.
Because the basal buried soil dates to 10 500 BP (TO-5280,
Table 2), while the Mazarna bed is at 2.2 m depth, the mean
rate of surface aggradation was 0.19 m.100 a-l between 10
500 and 6800 BP, and only 0.03 m. 100 a-I since 6800 BP. At
site 3 (Appendix A), Downie surface supports a low ridge
composed of 2.8 m of loess, including buried soils and a
Mazama tephra bed. The presence of loess on every

postglacial geomorphic surface from Maple Creek basin


shows that loess accumulation was a normal postglacial
process (Vreeken, 1993b).

INTERPRETATION
Characteristics of the geomorphic surfaces allow landscape
evolution to be reconstructed as a sequence of distinct phases.

Preglacial phases
The origin of Cypress Hills upland reflects late Tertiary fluvial responses to geological uplift, evidenced by the change
from Upper Cretaceous marine to Early Tertiary terrestrial
conditions. Fluvial aggradation formed Cypress Plain (Middle Miocene) and subaerial degradation, cyclically reactivated by base-level lowering, produced a series of Middle to
Late Miocene erosion surfaces. Late Miocene eolian deposition and minor fluvial action formed the Davis Creek silt, and
was followed by a prolonged sedimentary hiatus and soil formation. The product of these phases was a height of land that
imposed regional hydraulic gradients and a regionally anomalous
climate, influencing all subsequent geomorphic developments.

Glacial phases
Late Wisconsinan periglacial conditions began ca. 22 700 BP,
and frozen ground may have prevailed until maximally ca.
1 0 900 B P on surfaces that remained nonglaciated
(W.J. Vreeken, S.A. Wolfe, and D.J. Huntley, unpub. data,
1999). The Laurentide Ice Sheet reached the continental
divide ca. 22 000 BP, with Blacker Lake lobe (Hudson ice)
established prior to Gap lobe (Keewatin ice) (Klassen, 1994).
Proglacial meltwater eroded large channels across the divide
between these lobes and relicts of the Cypress Plain, with
smaller channels dissecting the eastern and central relicts
(Klassen, 1991, 1994; Vreeken, 1996b). Subglacial water
ascended the northern regional slope and incised a large channel, now bisecting the Gap lobe and occupied by upper Gap
Creek (1067-983 m; Fig. 3).
Deglaciation, beginning after ca. 18 000 BP, exposed the
highest driftplains and caused meltwater accumulation along
the northern slope with progressively lower eastward drainage, initially of the Carmichael meltwater system (approx.
1050-815 m a.s.1.; Klassen (1994)). Glacial Lake Downie,
the last phase of this system, was coeval with the 14 000 BP
active ice margin extending across the Triple Junction basin
(Fig. 12A). Exposure of the lakebed that forms the Downie
surface began prior to 13 100 BP, but lower parts were still
aggrading during a Glacier Peak tephra fall (ca. 12 000 BP),
after the active ice margin had shifted to the 13 000 BP Fox
Valley moraine. Northeast drainage of Lake Downie during
decay of the Triple Junction ice damlikely initiated formation
of lower Gap Creek valley. Exposure of Downie surface was
accompanied by melting and drainage of water from buried
ice, sediment surface collapse, sapping, and gravitational
failure of the walls of surface drainage-ways, causing significant but differential deglacial surface lowering.

.uo?s.ian?pp.lT)M]SOMi q ~ ~ M o ~ / o J M o ~ p.iZ)ML/$YIOS
z c / v ? ~Z/J?M
v ~ 'Wal~tk
~ ~ ~ sP ? J s ~ ? ~
JO( d g 000 $1 uvq1 ssal) sasuydisv1 puu aiv?pawta$u?(3 .'~oglv?2u18qns
pnvwsua puu xaldwo~Xallv~X O ayi
~ lv a2? an?iDuy q '231
~ an~ssud~ua~w~p
ayq
put7 aqol yi~o3u?x
8uyuodyaa~3a l d v a~ y q 1u?3ul8o~d
'walskspps8?gJo ( d g 000 EI(JOE EI '03)a s q d Xllva ( g .'aD? anti3v 8u?1uoAJa!u~oa
D aldvpq ayi u?uo?inlo~a
a 8 v u ? w ~ p ~ a i u ~1zn~u18aa
~ l a w ' ~a 1~ n z ? ~
1 ~ 3 v l 8 o ~L d~ a i ~ k ~ ~ ~a ~i D
u ~? /UmU( U
d g~000
J 91
O '03)a s q d ~ s v(?V : D ~ Ayaa~3

GSC Bulletin 534

The Bigstick meltwater system (approx. 800-715 m a.s.1.)


included an early phase with eastward flow, forming
east-trending glaciofluvial sediment ridges between the Triple Junction and Tenaille Lake areas (Fig. 12B), and was
likely sourced from the Kincorth moraine along the eastern
margin of Keewatin ice (Klassen, 1994). Glacial Lake Maple
Creek (800-765 m a.s.1.) began occupying the Triple Junction basin as it became ice free. While at approximately
800 m, glacial Lake Maple Creek received meltwater from
the south associated with melting of ice buried beneath the
Downie Lake plain, from the west while the ice front
remained at the Kincorth moraine (Klassen, 1994), and from
stagnant ice still covering lower terrain to the north
(Fig. 12B). Later, meltwater from the Bigstick Lake area
flowed in via a 38 km long, south-rising (715-765 m a.s.l.),
subglacial conduit that cut across and below the older
east-trending ridges, thereby initiating the formation of lower
Maple Creek valley (Fig. 12C). During the last Bigstick
phase, this southward flow was terminated by a westward
diversion through Bitter Lake meltwater channel near
Tenaille Lake (Fig. 12C). This diversion also completed formation of the Neitz surface and a scour basin below it which
became postglacial Bigstick Lake.

Postglacial fluvial processes operating within glacially and


deglacially formed basins were controlled by three distinct
base levels, which initially coexisted. The successive dominance of these base levels defined three fluvial phases. First
Junction Lake, then Hanson Lake, and ultimately modern
Bigstick Lake controlled upvalley stream erosion and
downvalley transmission of water and sediment (Fig. 13).
Thus, fluvial integration of the Maple Creek basin proceeded
stepwise, by amalgamation of downflow subbasins, without
significant headward channel extension. Eolian processes
acted during all fluvial phases.

. . .. ,
a

a.

MldHolocene

Early
Holocene

Late
Holocene

Climate
, , . . . ,

: . . .. .. . . . .

Eolian action

,,

, ,
',

Overland flow

Hvnrllh~rmal
.,--.

. . . , . . . . . ' .

. . , ,

Loess~deposits':,: : :,: : :,:

rn

L-1-1-1-1-1-kgstick

&:~ifier
Cakes,

:.,':1:1.;: :.I

Streamflow

\-

McDwgald channel

Glacial Lake Downie


(ca. 13 500 BP)

Site 5
Site 6
Site 7

I
I

b
w

Postglacial phases

Carbon
dates

Channel gradation

7601,

- --- - - - - -Base
1

16 000

- Junction Lake

Hanson Lake

, ,

720

levels

Glacial Lake Maple Creek

,
12 000

,-

= = =

' I

, , Lake, (max.
, 716
, m a.s.1.)
, , 1,
,
Bigstick
8000
Time (years BP)

4000

Figure 13. Chronological summary of postglacial climate,


loess accumulation, fan formation, and changes in the
fluviolacustrine system, base levels, and channel gradation in
Maple Creek basin. Abbreviations in climate intervals:
MWP, Medieval Warm Period; LIA, Little Ice Age.

Junction Lake base-level phase (ca. 12 000-6000 BP)


Junction Lake became the aggrading base level for Gap,
Maple, and McCoy basins when the diversion of meltwater
through Bitter Channel changed glacial Lake Maple Creek
into early postglacial Junction Lake (765 m a.s.1.). By
6800 BP, its maximum level was approximately 759 m a.s.1.
Ten metres of early postglacial incision of lower Gap
Creek, between 10 500 and 7300 BP at a mean rate of
0.31 m.100 a-l, is inferred as follows (Fig. 14). At site 7
(Appendix A), deoxidized or paleo-gleyed pond clay (81 1 m
a.s.1.) with shells dated at 11 800 BP (TO-5279) is overlain by
oxidized loess with a basal soil dated at 10 500 BP
(TO-5280). The change from wet to xeric site conditions,
between 11 800 and 10 500 BP, suggests drawdown of
initially high groundwater owing to incision of the adjacent
valley. A 7300-year old paleochannel beneath fan sediment at
site 6, about 1 rn higher than the modern channel bed, would
have been approximately 801 m a.s.1. near site 7 (Fig. 14).
Any channel bed, both upvalley from site 7 and less than
81 1 m a.s.l., should postdate 10 500 BP. This applies to the

buried channel at site 5 (approx. 808 m a.s.l.), which could


also be 7300 years old. This 'upvalley but lower than' principle for dating stream incision does not apply to the buried
channel at site 4, which is higher (approx. 822 m a.s.1.) and
may predate 10 500 BP.
The progress of valley gradation (Fig. 13) can be estimated, using supplemental information on deglacial lake levels. The highest hills on Downie surface are 838-876 m a.s.l.,
and occur within 1 km of the valley (Fig. 14). These were
exposed above lake level by 13 100 BP, since they are higher
than the site containing the youngest date for Downie Lake
shallow-water shells (site 1, Fig. 3; 831 m a.s.1.; Appendix A).
Hills above 861 m a.s.1. were emergent even earlier, when
glacial Lake Downie dropped below the elevation of
McDougald channel. Thus, local cross-valley topographic
relief produced during deglaciation was greater than the additional relief caused by stream incision since 10 500 BP. The
channel gradation curves (Fig. 13) suggest that deep, early
postglacial incision, in places down to bedrock, was

W.J. Vreeken

\,
I

Site 5

1. . . . ... . . .

Site 7

Site 6

. . ..

,,

McDougald
channel

..

-\\

Upland
elevations

7300 BP?

+I0500

Alluvial fans along lower Gap Creek formed during the


interval ca. 7300-5500 BP (Fig. 13). The high rate of fan
aggradation and associated sediment delivery to the creek are
consistent with the coeval, and possibly accelerated, construction of the Junction fan delta. Rapid aggradation in Junction Lake after 6800 BP, and consequent reduction of basin
volume, meant that decreased inflow was required to attain
outflow from the lake. In turn, this triggered accelerated retrograde incision of the Eagle outlet and increased the supply
of traction load to Hanson Lake, prior to final drainage of the
Junction Lake plain.

''

incision

Modern channel

7300BP

National Water Resources Research Institute, Service Report


88-0-219, 1988; Vreeken, 1993a, 1994). This increasing
aridity suggests that stream base flow originating from
groundwater discharge of better drained closed basins was
diminishing, resulting in a decreased inflow to Junction Lake
coincident with the occurrence of prolonged low water levels.

Distance (km)

Figure 14. Spatial and temporal relationships between


postglacial geomorphic su$aces from the lower reaches of
Gap Creek.

completed by ca. 8000 BP. Apart from temporary storage of


slump debris, lateral accretion, and fan deposits, only minor
aggradation has occurred since ca. 7300 BP.
Coarse (traction-load) sediment was largely trapped
within Junction and Hanson lakes, and is absent from
Bigstick basin clays older than 7500 BP. Until then, Bigstick
Lake extended well beyond its modern lake plain, indicating
that upstream lake basins were regularly filled to capacity.
The storage capacity of Bigstick Lake, perched on basal
postglacial clays, was gradually diminished by continued
sedimentation. Since there is no evidence that the lake ever
rose above adjacent Neitz surface (716 m a.s.l.), excess
inflow must have drained laterally into glaciofluvial deposits.
Accordingly, downbasin groundwater tables were influenced
by fluctuations of the upstream lakes. Clay beds in upper sediment sequences from the Bigstick Lake area (sites 18 and 19,
Appendix A) indicate reoccurrence of large lakes since
7500 BP. Because storage capacity was decreasing, these
occurrences required progressively smaller inflow. When
Hanson and Junction lakes were not filled to capacity, no
stream sediment would have reached the downstream basins.
Thus depositional hiatuses in the lower basins reflect, but also
exaggerate, the amplitude of hydrological variability in the
upper basin.
A hyperarid period with minimal interbasin flow,
between ca. 7500 and 6800 BP, is evidenced by the emersion
surface in Junction Lake, 4 m below outflow level. Severe
and/or prolonged aridity is also recorded by the thick overlying clay beds, which suggest that suspended sediment was not
being flushed from the lake basin. The timing of this interval
is consistent with ostracode data, from site 2 (Appendix A),
that record a change from parkland to prairie vegetation
between 8600 and 7000 BP (L.D. Delorme, unpub. data,

Hanson Lake base-level phase (ca. 6000-1500 BP,


Fig. 13)
With the demise of Junction Lake, coalescent Gap, McCoy,
and Maple creeks extended along a 9 km long trunk channel
to Hanson Lake, which provided a new local base level 27 m
lower than the former outlet of Junction Lake. After initial
headward channel incision, maximum geomorphic work was
accomplished during channel migration, beginning more
than 3600 BP. The pre-existing floodplain was consumed,
fans were truncated, valley-wall failure was reactivated, and
Weir surface was constructed. The effectiveness of Hanson Lake
in controlling water and sediment transmission is not known.
Bigstick Lake base-level phase (since ca. 1500 BP,
Fig. 13)
With the demise of Hanson Lake at ca. 1500 BP, base-level
lowering of about 33 m to Bigstick Lake added 25 km to the
uninterrupted master channel. The simultaneous addition of
the Bitter Lake distributary completed downflow basin arnalgamation. Subsequent headward channel incision extended
to the bedrock slopes of the upper Gap Creek basin, but was
followed by only minor channel migration.

DISCUSSION
Geomorphic controls on postglacial evolution
Maple Creek basin is the product of 10 to 15 million years of
landscape evolution typified by successive and interactive
fluvial, glacial, lacustrine, eolian, and mass-wasting systems.
These systems operated within and beyond the basin confines
at variably overlapping scales. Glacial deposition modified
an integrated preglacial fluvial basin, forming internally
drained hummocky topography and sediment barriers that
created nested subbasins. This imposed complex postglacial
hydrological connections between the climatically distinct
upper and lower parts of the basin. Deglacial meltwater
drainage evolution involved both proglacial and subglacial

GSC Bulletin 534

components, the former dominant during operation of the


Carmichael system and the latter during the Bigstick system.
Although the model of subglacial meltwater flow beneath
stagnant ice during formation of the Bigstick system is consistent with the observed relationships between landforms
and sediments, it and alternative explanations of proglacial
meltwater flow clearly require further examination.
The hummocky topography resisted reintegration of surface drainage by postglacial processes, and the cascade of fluvial subbasins broke down only about 1500 years ago. There
has been insufficient time for mutual adjustment of landforms
and postglacial stream processes. Thus, steep terrain with
outcropping bedrock in the upper basin still reflects
preglacial slope retreat followed by glacial erosion, while
today's valleys reflect their deglacial origin. Postglacial fluvial
action has barely extended beyond these preordained limits.

climate, ca. 6000 BP (Fig. 13), but it was the nonclimatic consequence of a lake reaching the limit of its sediment storage
capacity. However, the rate at which this limit was reached
was very likely climatically influenced. The transition to
Neoglacial conditions, the most humid late Holocene climate
phase (ca. 2700-1 100 BP; Vance et al. (1992)), may coincide
with the onset, before 3600 BP, of channel migration during
the Weir stream cycle (ca. 6000-1500 BP). klthough this
cycle overlaps with the Neoglacial, it primarily represents
fluvial adjustment to a new base level and is insufficiently
dated to permit additional climatic interpretation. Similarly,
the demise of Hanson Lake during latest Neoglacial times
was unrelated to ongoing climate changes. Geomorphic
responses that might coincide with the Medieval Warm
Period (ca. AD 900-1200; Hughes and Diaz (1994)) or with
the Little Ice Age (ca. AD 1450-1850; Luckman et al. (1993)) are
beyond the chronological resolution established for this study.

Climatic controls on postglacial evolution


Signals of regional climate changes from the geomorphic
record of Maple Creek basin are clouded by its steep climatic
gradient and by the many internal geomorphic controls discussed above. Although temporal coincidence between certain geomorphic and climate changes may suggest causality,
our understanding of basin history requires greatly increased
temporal resolution for such relationships to be assessed. The
fluvial record that overlaps with the relatively humid and cool
deglacial part of the Early Holocene climate interval (until ca.
9500 BP; Lemmen et al. (1998)) was dominated by stream
incision to the Junction Lake base level (Fig. 13). However,
because this was not accompanied by channel network
development, it represents a delayed response to deglaciation
that was favoured, but not controlled, by the prevailing
regional climate.
The onset of warmer and &ier early Holocene conditions
may have been diachronous within the basin. The change
from parkland to prairie on the continental divide, between
8600 and 7000 BP (site 2, Appendix A), occurred later than a
similar change at Harris Lake (ca. 9000 BP; Fig 1B; Sauchyn
(1990)), but coincided with the onset of Hypsithermal conditions (Fig. 13), recorded as the warmest and driest between
7700 and 6000 BP at Chappice Lake (Fig. 1B; Vance et al.,
1992). The latter site is similar to the Bigstick Lake area
where, nonetheless, a vast lake prevailed during the preceding interval of increasing aridity (until ca. 7500 BP). Disappearance of this lake would suggest that a basin-wide
hydrological threshold was exceeded ca. 7500 BP, and is
attributed to a reduced moderation exerted by the upland climate. Junction Lake may have been temporarily dry between
7500 and 6800 BP, but much of the time it represented an
important water body, maybe regionally unique, because it
was maintained by the upland climate. Geomorphic
responses that are better correlated, and more consistent, with
this interval of maximum aridity include the formation of
alluvial fans (7500-5500 BP), associated increased sand supply from flash floods into Junction Lake, and cyclic
high-intensity slope erosion in closed small basins (Ham site,
Fig. 1B; Vreeken (1996a)). The demise of Junction Lake may
have coincided with a change to less arid mid-Holocene

ACKNOWLEDGMENTS
This study was funded by the Natural Sciences and Engineering Research Council of Canada (NSERC Grant A3872) and
from a Palliser Grant by Natural Resources Canada. Early
parts of the fieldwork were done in collaboration with
R.W. Klassen (Geological Survey of Canada). Jeff Bond
assisted during coring operations. Janis Dale (University of
Regina) analyzed mollusc shells. Prairie Farm Rehabilitation
Administration (PFRA) personnel at Maple Creek provided
useful documentation. Eric Lawrence, Ted Friday,
Mary-Anne Friday, and many other landowners in the area
extended co-operation and kindness. I thank D.S. Lernmen
(Geological Survey of Canada) for his assistance during the
writing of this paper.

REFERENCES
Adon, D.F., Clayton, J.S., Ellis, J.G.,
Christiansen, E.A., and Kupsch, W.O.
1960: Physiographic divisions of Saskatchewan; Saskatchewan Research
Council, Geology Division, Map 1, scale 1:520 640.
Alden, W.C.
1932: Physiography and glacial geology of eastern Montana and adjacent
areas; United States Geological Survey, Professional Paper 174,
133 p.
Allen, J.R.L.
1984: Sedimentary structures: their character and physical basis; Developments in Sedimentology, v. 30, Elsevier, Amsterdam, 663 p.
Bacon, C.R.
1983: Eruptive history of Mount Mazama and Crater Lake caldera, Cascade Range, U.S.A.; Journal of Volcanology and Geothermal
Research, v. 18, p. 57-1 15.
Barendregt. R.W.. Vreeken, W.J..
Irving, E.. and Baker, J.
e
Creek
1997: S&tigraphy and paleomagnetismldf ththk ate ~ i o c e n Davis
silt. east block of the Cv~ressHills. Saskatchewan: Canadian Journal'of Earth Sciences, 34, p. 1325-1332.
Canada Centre for Mapping
1990: Canada: climatic regions, Thomthwaite classification [map]: moisture regions; Map 4.12 in National Atlas of Canada; Energy, Mines
and Resources Canada, Geographical Services Division, scale
1:7 500 000.
Canadian Agricultural Services Coordinating Committee
1983: The Canada Soil Information System (CanSIS). Manual for
Describing Soils in the Field; Agriculture Canada, Research
Branch, Ottawa, Ontario, 97 p. (plus five appendices).

;:

W.J. Vreeken

Christiansen, E.A.
1979: The Wisconsinan deglaciation of southern Saskatchewan and adjacent areas; Canadian Journal of Earth Sciences, v. 16, p. 913-938.
Clayton, J.S., Ehrlich, W.A., Cann, D.B., Day, J.H., and Marshall, I.
1977: Soils of Canada: Volume 1, Soil Report; Canada Department of
Agriculture, Research Branch, Ottawa, Ontario, 243 p.
David, P.P.
1964: Surficial geology and ground water resources of the Prelate area
(72K). Saskatchewan; Ph.D, thesis, McGill University, Montrial,
Quebec, 329 p.
Dyke, A.S. and Prest, V.K.
1987: Late Wisconsinan and Holocene history of theLaurentide Ice Sheet;
GCographie physique et Quaternaire, v. 41, p. 237-263.
Eyles, N., Eyles, C.H., and Miall, A.D.
1983: Lithofacies types and vertical profile models: an alternative
approach to the description and environmentalinterpretationof glacial diamict and diamictite sequences; Sedimentology, v. 30,
p. 393410.
Foit, F.F., Mehringer, P.J., and Sheppard, J.C.
1993: Age, distribution,and stratigraphyof Glacier Peak tephra in eastern
Washington and western Montana; Canadian Journal of Earth Sciences, v. 30, p. 535-552.
Hughes, M.K. and Diaz, H.
1994: Was there a 'Medieval Warm Period', and if so where and when?;
Climatic Change, v. 26, p. 109-107.
Klassen, R.W.
1991: Surficial geology and drift thickness, Cypress Lake, Saskatchewan;
Geological Survey of Canada, Map 1766A, scale 1:250 000.
1992: Nature, origin, and age relationships of landscape complexes in
southwestern Saskatchewan; GBographie physique et Quaternaire,
v. 46, p. 361-388.
1994: Late Wisconsinan and Holocene history of southwestern
Saskatchewan; Canadian Journal of Earth Sciences, v. 31,
p. 1822-1837.
Kor, P.S.G., Shaw, J., and Sharpe, D.R.
1991: Erosion of bedrock by subglacial meltwater, Georgian Bay,
Ontario: aregional view; Canadian Journal of Earth Sciences, v. 28,
p. 623-642.
Leckie, D.A. and Cheel, R.J.
1989: The Cypress Hills Formation (Upper Eocene to Miocene): a
semi-arid braidplain deposit resulting from intrusive uplift;
Canadian Journal of Earth Sciences, v. 26, p. 1918-1931.
Lemmen, D.S., Vance, R.E., Campbell, I.A., David, P.P.,
Pennock, D.J., Sauchyn, D.J., and Wolfe, S.A.
1998: Geomorphic systems of the Palliser Triangle, southern Canadian
Prairies: description and response to changing climate; Geological
Survey of Canada, Bulletin 521,72 p.
Luckman, B.H., Holdsworth, G., and Osborn, G.D.
1993: Neoglacial glacier fluctuations in the Canadian Rockies; Quaternary Research, v. 39, p. 144-153.
Miall, A.D.
1978: Facies types and vertical profile models in braided river deposits: a
summary; in Fluvial sedimentology, (ed.) A.D. Miall; Canadian
Society of Petroleum Geologists, Memoir 5, p. 597-604.
1992: Alluvial deposits;in Facies Models: Response to SeaLevel Change,
(ed.) R.G. Walker and N.P. James; Geological Association of
Canada, p. 119-142.
Ruhe, R.V.
1975: Geoinorphology: Geomorphic Processes and Surficial Geology;
Houghton Mifflin Company, Boston, Massachusetts, 246 p.

Saskatchewan Institute of Pedology


1990: Preliminary soil map and report of Big Stick Rural Municipality no.
141, Saskatchewan; Saskatchewan Institute of Pedology, University of Saskatchewan, Saskatoon, Saskatchewan, 27 p. (map at
1:100 000 scale).
Saskatchewan Institute of Pedology (cont.)
1991: The soils of Maple Creek Rural Municipality no. 11 1,
Saskatchewan; Saskatchewan Institute of Pedology, University of
Saskatchewan, Saskatoon, Saskatchewan, Publication SM I I I,
66 p. (map at 1:100 000 scale).
Saskatchewan Research Council
1987: Surficial geology of the Prelate area (72K), Saskatchewan;
Saskatchewan Research Council, Saskatoon, Saskatchewan, scale
1:250 000.
Sauchyn, D.J.
1990: A reconstruction of Holocene geomorphologyand climate, western
Cypress Hills, Alberta and Saskatchewan; Canadian Journal of
Earth Sciences, v. 27, p. 1504-1510.
Storer, J.E.
1975: Middle Miocene mammals from the Cypress Hills, Canada;
Canadian Journal of Earth Sciences, v. 12, p. 52CL522.
Vance, R.E., Mathewes, R.W., and Clague, J.J.
1992: A 7000-year record of lake-level change on the northern Great
Plains: a high resolution proxy of past climate; Geology, v. 20,
p. 879-882.
Vreeken, W.J.
1984: Relative dating of soils and paleosols; in Quaternary Dating
Methods, (ed.) W.C. Mahaney; Elsevier Science Publishers,
Amsterdam, The Netherlands, p. 269-281.
1986: Quaternary events in the Elkwater Lake area of southeastern
Alberta; Canadian Journal of Earth Sciences, v. 23, p. 2024-2038.
1993a: Postglacial soil-landscape regimens in the Palliser Triangle; in The
Palliser Triangle: A Region in Space and Time, (ed.)
R.W. Barendregt, M.C. Wilson, and F.J. Jankunis; Thompson's
Communications,Calgary, Alberta, p.125-152.
199323: Loess and paleosols in southwestern Saskatchewan and southern
Alberta; in Quaternary and Late Tertiary Landscapes of Southwestern Saskatchewan and Adjacent Areas, (ed.) D.J. Sauchyn;
Canadian Plains Research Centre, University of Regina, Regina,
Saskatchewan,p. 2 7 4 5 .
1994: A Holocene soil-geomorphic record from the Ham site near Frontier, southwestern Saskatchewan; Canadian Journal of Earth Sciences, v. 31, p. 532-543.
1996a: A chronogram for postglacial soil landscape change from the
Palliser Triangle, Canada; The Holocene, v. 6, p. 433438.
1996b: Belanger canal, Cypress Hills, in Landscapes of the Palliser Triangle, (ed.) D.S. Lemmen; Canadian Association of Geographers,
Field Guide, p. 54-55.
Vreeken, W.J. and Westgate, J.A.
1992: Miocene tephra beds in the Cypress Hills of Saskatchewan,Canada;
Canadian Journal of Earth Sciences, v. 29, p. 48-51.
Vreeken, W.J., Klassen, R.W., and Barendregt, R.W.
1989: Davis Creek silt, anEarly Pleistoceneor Late Pliocenedeposit in the
Cypress Hills of Saskatchewan; Canadian Journal of Earth Sciences, v. 26, p. 192-198.
Vreeken, W.J., Westgate, J.A., and Alloway, B.V.
1992: Geomorphic significance of Miocene rhyolitic tephra beds from the
Cypress Hills, Saskatchewan, Canada; Quaternary International,
v. 13/14, p. 23-28.
Westgate, J.A. and Naeser, N.D.
1985: Tephrochronology and fission-track dating; in Dating Methods of
Pleistocene Deposits and Their Problems, (ed.) N.W. Rutter; Geoscience Canada, Reprint Series 2, p. 31-38.
Young, R.R., Burns, J.A., Smith, D.G., Arnold, L.D., and Rains, R.B.
1994: A single, late Wisconsinan, Laurentide glaciation, Edmonton area
and southwestern Alberta; Geology, v. 22, p. 683-686.

GSC Bulletin 534

APPENDIX A
Site descriptions
See Figures 3,6,7,8, and 9 for locations of the sites described here. Radiocarbon dates are presented in Table 2.

Common legend for sections


Facies code ~ithofacies'

Interpretation

Fluvial and lacustrine facies:


Gm

Gravel: sorted; massive or crudely bedded (horizontal imbrication)

Long bars; lag or sieve deposit

Gms

Gravel: matrix supported; massive

Debris-flow deposit

Sh

Sand: very fine to very coarse, may be pebbly; horizontal laminae

Parting or streaming lineation; planar bed flow


(upper flow regime)

SP

Sand: medium to very coarse, may be pebbly; solitary or grouped


planar crossbeds

Linguoid, transverse bars, sand waves (lower flow


regime)

Sr

Sand: very fine to coarse; ripple crosslaminae

Ripples (lower flow regime)

Ss

Sand: fine to very coarse, may be pebbly

Shallow scour fills

St

Sand: medium to very coarse, may be pebbly

Solitary or grouped trough crossbeds; dunes


(lower flow regime)

Fc

Clay: laminated or apparently massive

Lacustrine suspension deposit

Fcgm

Clay: gritty, poorly sorted; massive

Sediment gravity-flow deposit

Fh

Loam (clay, silt, and sandy; rare discontinuous clay laminae; faint
horizontal bedding planes

Subaerial alluvial fan deposit

FI

Sand, silt, and clay: fine laminae, very small ripples

Overbank or waning flood deposit

Fm

Clay and silt: massive, often with desiccation cracks

Overbank or drape deposits

Glacial, glaciofluvial, and glaciolacustrine facles:


Dmm

Diamict: matrix supported; massive

Till

Smd

Sand: medium to coarse; sorted, rare dropstones; massive, locally


weakly graded

Turbidite(?) with minor ice-rafted debris

Fld

Silt or fine sand, and clay: rare dropstones; rhythmically laminated

Suspension deposits with minor ice-rafted debris

Eolian facies (e):


Fme

Loam (clay, silt, and sandy): massive

Loess

Sme

Sand: sorted; massive

Sandy loess

Silt (volcanic glass): sorted; massive

Tephra

. .

Buried soils
Control elevation

- fl
- Radiocarbon date
DU Deoxidized unleached sequencesz

Texture (gravel I sand I s~lt& clay)

Laminated sediments

Lithofacies notations after Miall (1978, 1992) and Eyles et al. (1983).
In deoxidized material, 60% of the matrix has MunsellTMhues of IOYR, 2.5Y, and/or 5Y, values of 5 and 6, and chromas of 1 or 2 with
segregation of iron (ferric oxides) and manganese into tubules (pipestems) or nodules (Hallberg et al., 1978). The light grey matrix has
low total free iron and very low reduced iron values. This shows the effects of mobilization and movement of iron compounds (gleying),
under conditions of prolonged high water saturation, as with a high groundwater table. When such iron segregation is encountered in
presently well drained materials, it is interpreted as a relict condition caused by once higher water tables (Hallberg et al., 1978).

W.J. Vreeken

South

SITE 1 - Fleming Creek site (Fig. 3)


Geomorphic surface: Downie surface
Latitude: 49'49'20"N
Longitude: 109"29'00" W
Su@ce elevation: 833 m a.s.1.
Description: Roadcut through hilltop on hummocky lake
plain (815-907 m a.s.1.) exposes approximately 3 m of
inclined and normal-faulted, planar crossbedded silt and
sand. Two metres below the surface, a 0.3 m thick loam
diamict bed contained Stagnicola shells dating 14 000 BP
(GSC-4675, conventional) and 13 100 BP (TO-694, AMS).
Similar deformed sand and silt beds are exposed around
nearby Harris Reservoir.

North

Interpretation: The sequence accumulated in glacial Lake


Downie. Because the shells are from a shallow-water gastropod, the lake level was at most a few metres above 833 m
a.s.1. when they were deposited. Sediment deformation
reflects collapse upon melting of buried glacier ice.
Significance: Sediments accumulated shortly before the
demise of glacial Lake Downie, during the last phase of the
Carmichael meltwater drainage system (approximately
1050-815 m a.s.1.; Klassen (1994)). That system discharged
east, along the north flank of Cypress Hills upland, and ultimately along the active ice margin that extended across the
location of the town of Maple Creek (Vreeken, 1989;
Klassen, 1994). The shells and a Glacier Peak tcphra bed (at
site 7) in the uppermost sediments indicate open-water conditions. Thus, by about 13 100 BP, terrain south of site 1 and
above approximately 833 m a.s.1. was subaerially exposed.
Melt-out resulting in sediment deformation continued after
this date.

st
1

SITE 2 - Black site (Fig. 3, A-1)


Geomorphic su@ce: Upper Gap surface
Latitude: 49'37' 10" N
Longitude: 109"42'30"W
Su$ace elevation: 11 12 m a.s.1.
Description: Boreholes at the bottom of a closed basin in a ridged,
hummocky surface on the continental divide (Vreeken, 1993a) terminate in sandy loam diamict, interbedded with sorted sand and
clast-supported gravel. Above this lies 5.5 m of interbedded clay, silt
loam, and sandy loam that includes a Mazama tephra bed and organically enriched buried soils. Shells from snails and bivalves at 3.2 m
depth date to 8600BP (TO-201). Ostracode assemblages are present
from 4.2 to 2.5 m depth. A surficial mantle (0.3 m) of massive sandy
loam extends to the basin perimeter where it rests on a buried soil
formed from diamict.
Interpretation: The substrateconsists of sedimentgravity-flow and
current-bedded deposits that may have accumulated subglacially.
The overlying sediments reflect a change from perennial to ephemeral pond conditions after ca. 8600 BP. Ostracode assemblages

Stoneline
Shells
Paleosol
14csample

Figure A-I.
Borehole stratigraphy of enclosed basin at site 2
(modified from Vreeken, 1993~).
reflect a vegetation change from parkland to prairie shortly before
their disappearance, ca. 7000 BP (L.D. Delorme, unpub. data,
National Water Resources Research Institute, Service Report
88-0-219, 1988; Vreeken, 1993a, 1994, 1996a). Locally confined
clay reflects slope wash within the closed basin, while buried soils
indicate that erosion was cyclic. The surficial sandy loam is eolian.
Significance: Wet site conditions ended ca. 8600 BP, slightly after
regional onset of the Hypsithermal climate interval (ca. 90004500 BP). The transition from parkland to prairie correlates with the
onset of warmest and driest Hypsithermal conditions (77006000 BP; Vance et al. (1992)). Aridification on the continental
divide may have lagged behind regional climate change.

GSC Bulletin 534

SITE 3 - Downie Lake site (Fig.3, A-2)


Geomorphic su$ace: Downie surface
Latitude: 49'47'50"N
Longitude: 109'40'30"W
Surj5ace elevation: 884 m a.s.1.

SITE 4 - Southwood's fan site (Fig.9, A-3)


Geomorphic su$ace: Lawrence surface complex
Latitude: 49'49' 10"N
Longitude: 109'34'55"W
Surface elevation: approximately 828 m a.s.1.

Description: Bluff along east shore of Downie Lake cuts across a


low east-trending ridge on the nondissected Downie surface. Basal
normal-faulted and slightly inclined, parallel-laminated and
crossbedded, deoxidized medium sand (Sp, DU) is truncated by a
horizontal erosional surface. The 2.8 m thick overburden consists of
upward-convex lenticular units of massive oxidized loamy sand and
sandy loam (Fme), separated by buried soil mantles and a Mazama
tephra bed (T). The basal buried soil has a thick black decalcified A
horizon, but the other buried soils are dark greyish brown and
calcareous.

Description: Cutbank along Gap Creek (approximately 821 m


a.s.1.) exposes the substrate of an alluvial fan that descends from
Downie surface. The truncated fan toe stands above a Weir surface
terrace. Basal deoxidized silt-clay rhythmites (Fld, DU) are
erosionally overlain by deoxidized matrix-supported sandy gravel
(Gm, DU) with lenticular and planar sand interbeds (St, Sh). Clast
(including Precambrian Shield rock types)
imbrication and crossbedding record
0downvalley flow. The gravel is overlain by
oxidized sand-clay rhythmites (Fl), with
crosslaminated sands dipping downvalley.
1The overlying sediments are a complex of
oxidized, generally massive sandy loam
(Fme), intercalated with thin sand beds
(Sme), a tephra bed (T), and several buried
pL
soils. The lowest soil has a 40 cm thick black
decalcified Ah horizon, while the others are
greyish brown and calcareous. The upper- ,
most sandy loam (Fh) is faintly bedded,
3-includes rare clay laminae, and extends to the 5
fan surface.
0"

Fme

*884

a.s'''

Figure A-2.
section, site 3.

Interpretation: The basal sediments accumulated in glacial Lake Downie (see site 1) and
were deformed by the melting of buried ice.
Subsequent planation occul~edduring drainage of the Downie Lake plain. Postglacial soil
formation alternated with eolian accumulation
of material from a nearby source. The structure of the eolian layers suggests that the ridge
is an accretionary feature and its orientation is
consistent with postglacial prevalence of
westerly winds. Colour and thickness of the
basal paleosol suggest that it formed in a moist
site environment, which is consistent with the
deoxidized state of the permeable sand substrate. R a i n f a l l was adequate f o r
decalcification. Weak development of the
younger buried soils may reflect less time for
formation, combined with increased aridity.

Significance: Cyclical accumulation of sandy


loess began in early postglacial times, during a
relatively moist and cool climate. Deposition continued throughout
the warm and dry Hypsithermal, producing the constructional
wind-aligned ridge.

:
:

Interpretation: The basal rhythmites accumulated in glacial Lake Downie, but the overlying gravels are likely from a postglacial
channel. The next rhythmite sequence
denotes waning flow and channel diversion,
possibly caused by slumping of the valley
wall. The massive sandy loam deposits are
loess and the tephra bed is probably Mazama,
as at site 3. The morphology of the basal
paleosol suggests formation under moist conditions, whereas the soils above it reflect
either or both of shorter soil-forming intervals and xeric site conditions. The uppermost
bedded loam is directly associated with the
fan surface and must be a fan deposit.

4-

5-

6-

* 822 m a.s.1.
Figure A-3.
Stratigraphic
section, site 4.

Significance: This site, together with sites 5


and 6, documents the Lawrence alluvial fan substrates. The thick
loess with its prominent basal soil suggests that the channel deposit
is much older than the tephra bed (6800 BP). Thus, the channel and
overlying rhythmites, hence the inferred slump event, may date from
the early postglacial. The fan was completed less than 6800 BP but
before completion of the Weir stream cycle.

W.J. Vreeken

SITE 5 - Lawrence's fan site (Fig.9, A-4)


Geomorphic suvace: Lawrence surface complex
Latitude: 49'49'55"N
Longitude: 10933'15'fW
Surface elevation: approximately 814 m a.s.1.

SITE 6 - Gap Creek 2 site (Fig. 9, A-5)


Geomorphic surface: Lawrence surface complex
Latitude: 49'5 1'05"N
Longitude: 109O35'25"W
Surface elevation: 803 m a.s.1.

Description: Cutbank along Gap Creek (approximately 808 in


a.s.1.) shows the substrate of an alluvial fan that descends from
Downle surface. The upper fan IS partly separated from the upland
by the headscarp of a major slump. Basal sediments consist of
clast-supported gravel (Gm) with matrix-supported gravel interbeds
(Gms). This unit is overlain by a complex of thinly interbedded sedlments, coinprlsing very thin, sorted, medium
sand (Sh), apparently massive, and massive,
gritty to pebbly, clay loam diamict beds
(Fcgm) that plnch Into short (0.5 m) pebble
bands. This sequence is capped by clay (Fc). A
tephra bed (T) on the clay 1s overlain by massive sandy loam (Fme), and capped by crudely
bedded sandv loam (Fh) that extends to the fan
21
surface.

Description: Cutbank along Gap Creek (approximately 792 m


a.s.1.) shows the substrate of an alluvial fan that descends from
Downie surface. The upper fan is partly separated from the upland
by the headscarp of a major slump. The truncated fan toe stands
above a terrace of the Weir surface and an alluvial bench of the Bull
surface. Exposure reveals Bearpaw shale (bedrock) overlain by a
1.1 m thick deoxidized complex consisting of ripple- and planarlaminated fine sand (Sr), with the upper scours filled by laminated
coarse to medium sand (Ss) with wood fragments, and capped by
veiy thin, coarse sand and s ~ lrhythmltes
t
(Fl) with wood and bark
fragments. Salix fragments from the latter
were dated at 7300 BP (GSC-4422 and
GSC-6068). Overlying the deoxidized unit are
oxidized sediments fining upwards from
gravel (Gm) to sand (St), clay and silt (Fm),
and capped by a very thick (7.3 m) unit of oxidized, parallel thin- and medium-bedded, fine
sandy and silt loam (Fh). A Mazama tephra
bed occurs 5.3 m below the surface.

Interpretation: Postglacial channel flow was


initially impeded and then diverted, probably
due to slumping of the valley wall. The interbedded complex represents sediment gravity
flows into a body of through-flowing water.
Deposition of clay from suspension occurred
after flow stagnated. The tephra bed, likely
Mazama (6800 BP), and the overlying loam
accumulated subaerially, initially as loess
(Fme) and subsequently as fan deposits (Fh).

Figure A-4.
Stratigraphic
section, site 5 .

Significance: This site, together with sites 4


and 6, documents Lawrence alluvial fan substrates. The buried channel deposits predate
6800 BP. Their maximum age is less than
10 500 BP, based on the elevation of a dated
sample downstream (site 7). Subsequent
slumping of the valley wall occurred prior to
6800 BP, and was followed by loess accumulation and aggradation of the fan.

A shallow section through the Bull surface


bench, 2 m above Gap Creek, shows 1 m of
planar crossbedded sand separated by a
weakly developed buried soil from overlying
massive clay (0.3 m). The surface of the buried
soil is marked by scattered bones and a hearth
with charred wood dated 40 k 5 0 BP
(GSC-4421).
Interpretation: The basal deoxidized sand is
a splay deposit, whereas the overlying oxidized gravel and sand are lateral accretion
deposits capped by overbank deposits. The
remainder of the section records fan
aggradation. The section in the Bull surface
bench is composed of lateral accretion
deposits capped by overbank clay.
Significance: This site, together with sites 4
and 5, documents Lawrence alluvial fan substrates. Postglacial stream incision to bedrock
occurred before 7300 BP and was followed by
minor stream aggradation. Fan formation
began between 7300 and 6800 BP, and terminated after 6800 BP. The Bull surface bench is
modern.

rock

* 793 m a.s.1.
7270 i 80
(GSC-4422)
7250 i 100
(GSC-6068)

Figure A-5.
Stratigraphic
section, site 6.

GSC Bulletin 534

SITE 7 - Friday's site (Fig. 9, A-6)


Geomorplzic sur$ace: Downie surface
Latitude: 49'5 1'00"N
Longitude: 10933'55" W
Surface elevation: 823 m a.s.1.

SITE 8 - Weir radiocarbon site (Fig. 9, A-7, A-8)


Geomorphic surface: Weir surface
Latitude: 49'5 1'33"N
Longitude: 109O35'0" W
Surface elevation: 785 m a.s.1.

Description: Closed-basin fill is exposed in a


slump scarp on a high-relief part of Downie surface, immediately above a Weir surface terrace.
Thin tilted beds of deoxidized planar and
Frne
low-angle crosslarninated, medium and very
1fine sand (Sp and Sr) at the base of the exposure
include a 0.3 cm thick Glacier Peak tephra bed,
and are overlain by a 0.5 m thick clay bed (Fc)
21-T with abundant Stag~zicolashells dated 11 800
BP (TO-5279). These sediments are overlain
by 9.2 m of oxidized loam deposits that include
40 buried soils and a Mazama tephra bed at
32.1 m depth. The loam ranges from sandy to silt
E - > ,
and is generally massive (Frne). Many of the
- <
5
a Fme
buried soils contain charcoal, with fragments
from the lowermost buried soil (811 m a.s.1.)
d 8dated 10 500 BP (TO-5280).

Description: Cutbank along Gap Creek


*.::.....
. .
(778 m a.s.1.) exposes substrates of Weir terrace (785 m a.s.1.) and Downie surface (823
- ...... .. Fm
m a.s.1.; Fig. A-8). Loam diamicton (Drnm)
- . .
with striated boulders underlies both fea1 - . : ',
tures. Beneath the upland, this is overlain by
approximately 3 8 m of horizontal,
- . .
. S r
medium-bedded sand and silt. Beneath the
- .':
terrace, the top of the diamicton is
E 2 .: .
erosionally overlain by 0.3 m of scattered
- ' : '.
boulders and crudely bedded, deoxidized 'L
sandy pebble gravel (Gms) with rare bivalve 8
.'.
shells and reworked bone fragments, which
in turn is overlain by approximately 3.2 m of
fining upward sand (from deoxidized, planar crossbedded, and very coarse (Sp) to
oxidized, subhorizontal, faintly thinbedded, and fine (Sr)), with abundant
Sphaeriurn shells, and capped by 0.5 m of
massive clay (Fm) and discontinuous mas*785 m a.s.1.
sive sandy loam (Sme). A bone from thepla2940 + 160
nar crossbedded sand was dated 3600 BP
(GSC-4027)
0 3600 + 80
(TO-loo), while shells from the overlying
(TO-I 00)
fine sand were dated 2900 BP (GSC-4027).
At a nearby site, a bone from a correlative
capping clay was dated 1500BP (TO-4416).
~i~~~~
A-7.

0-

--

Interpretation: The basal sand was deposited


in glacial Lake Downie, with the hummocky
topography and sediment deformation
reflecting collapse during melting of buried ice.
The deoxidized clay bed was deposited in an
early postglacial pond, indicating a high
groundwater table. The overlying loam is dominantly loess that accumulated cyclically in a
xeric site environment prone to fire. Occasional
graded bedding reflects variations in wind
strength and local reworking by slopewash.
The slump scarp predates the Weir terrace.

Figure A-6.
Stratigraphic
section, site 7.

Significance: Downie surface was subaerially


exposed and its hummocky topography formed
shortly after the Glacier Peak eruption. Loess
accumulation began between 11 800 and
10 500 BP following groundwater table
drawdown below 81 1 m a.s.1. Drawdown likely
relates to incision of Gap Creek, which
therefore was less than 811 m a.s.1. bv
10 500 BP.

Figure A-8. Geomorphic surfaces and cutbank


exposure of site 8, looking north. Northernmost part of
Downie Lake plain is visible on the upland, with the
Triple Junction lowland in thefar distance. Photograph by
W.J. Vreeken. GSC 1999-0430

.I

: :

~trktigra~hic

Interpretation: The upland section consists


section, site 8.
of till overlain bv.de~osits
- from glacial Lake
Downie. The terrace section records stream
incision before 3600 BP and subsequent aggradation of channel-bed
gravel deposits, lateral accretion deposits, overbank clay, and loess.

Significance: The fluvial sediments date the Weir stream cycle


between ca. 3600 and 1500 BP.

W.J. Vreeken

SITE 9 - fan delta site (Fig. 6, A-9)


Geomorphic suvface: fan delta of Junction surface complex
Latitude: 49'52'25"N
Longitude: 109'34'50" W
Su$ace elevation: 777 m a.s.1.

SITE 10 - Mouse Creek site (Fig. 6, A-10)


Geomoiphic surface: lake plain of Junction surface complex
Latitude: 49'55'55"N
Longitude: 109"00'30" W
Su$ace elevation: 758 m a.s.1.

Description: Large cutbank into a surface that slopes and broadens


from Gap Creek valley into Triple Junction basin exposes numerous
crosscutting channel scours that are dominantly filled by
scour-conform, thin- and medium-bedded, planar-laminated,
medium and fine sand (Sh). Uppermost fills are commonly capped
by thin-bedded, laminar couplets of fine sand and clay (Fl). Weakly
developed soil A horizons commonly surmount the clay beds.

Description: A 2.5 m high cutbank along a small creek, incised


below Junction Lake plain, is connected to 3 4 m high continuous
cliffs along the south shore of Junction Reservoir. Site is at the west
end of this approximately 2.5 km long exposure. Sediments show
alternation of thick sand-dominated units with thick clay beds. To
facilitate identification and tracing, each sand and clay unit was designated alphabetically (Fig. A-1OA). The sand-dominated units are
composites of medium- to thin-bedded, fine and medium sand (designated with suffix s), interbedded with thin to very thin clay (designated c), and numbered sequentially within each cycle. The thick
clay beds are denoted 'cx'. One thick clay bed contains a slumped
tephra bed, identified as Mazama and found only at site 10
(Fig. A-1OB). This key bed is named 'Mcx' and all further designations follow from this.

Interpretation: Scour-and-fill structures were produced by


abruptly shifting channels during high stream and sediment discharge events. Cyclical aggradation of sand and clay couplets
occurred during minor intervening floods of revegetated surfaces.
Significance: The scour-and-fill sequences are consistent with the
debouchure of a relatively narrow valley into a broad basin, recording development of a fan delta. The abrupt super- and juxtaposition
of the scours suggests that they resulted from recurrent flash floods,
possibly reflecting an arid climate with high-magnitude precipitation events.

The cx clay beds (Fc) are very dark greyish brown and commonly
massive, although the thin clay layers are usually faintly laminated.
Sand beds consist of thin tabular sets of cross- and planar-laminated,
fine and very fine, pale brown oxidized sand (Sr). The Ocx bed is
overlain by 0.2 m of massive sandy loam. See site 11 for further
descriptions.

A minor channel fill, incised 2 m below the lake plain and located
approximately 20 m north of site 10, contained a large bone that was
dated 2300 BP (TO-4415).
Interpretation: Each unit of sand and thick silt is interpreted to represent a major cycle of lacustrine sedimentation in Junction Lake.
TheM, N, and 0 cycles are recorded at this site. Thechannel with the
dated bone postdates drainage of Junction Lake, while surficial
sandy loam is interpreted as loess.
Significance: Junction Lake existed before and after the Mazama
eruption (6800 BP), during the warmest and driest part of the
Hypsithermal climate interval, a time when most lakes in the Palliser
region had dried up. The 2300-year old channel fill accumulated
during the Weir depositional cycle (3600-1500 BP; see site 8),
which postdates Junction Lake.

Figure A-9. Crosscutting scour channels of Junction fan


delta. Fill sequences numberedfrom oldest to youngest. Photograph by W.J. Vreeken. GSC 1999-043E

Figure A-10. Junction


Lake sediments at site
10: A) cyclically alternating fine sand and
clay (see text for discussion); scale is 2 m,
GSC 1999-043F; B)
slunzped Mazama
tephra, intercalated
within unit Mcx; scale
is 0.6 rn. GSC 1999043G. Photographs by
W.J. Vreeken.

GSC Bulletin 534

SITE 11 - Yarshenko's transect (Fig. 6, A-1 1, A-12)


Geornorphic surjface: lake plain of Junction surface complex

Latitude: 49'56'16"N
Longitude: 109"29'30"W
Surjface elevation: approximately 757 m a.s.1.
Description: This transect, at the east end of continuous cliffs along
the south shore of Junction Reservoir, records sedimentation cycles
K, L, M, N, and 0 of Junction Lake (see site 10) and intraformational
channels, not observed elsewhere. Thick cx clay beds (Fc) are as
described for site 10, but the Ncx bed includes a discontinuous, thin,
dark band of fine-grained charcoal or lignite and very rare shells
between 0 and 20 m (Fig. A-l 1). The Ocx clay bed thins to zero
between sites 10 and 11, such that surficial sandy loam rests directly
on the Osc complex. The Lsc and Msc units are absent, while the Nsc
unit wedges out. Crosslaminated sand (Sr) from the Nsc and Osc
units has north-northeast-dipping foresets. Transitions from rippled
sand to clay are usually abrupt and planar, and interlamination was
seen in only a few places. Transitions from massive Mcx and Ncx
clay to sand are also abrupt, but planar to wavy and, in places, convoluted. No infilled desiccation or syneresis cracks extend below these
contacts, nor do clay-balls (reworked mud) occur in the overlying
rippled-sand facies. The Kcx bed is unusual, featuring many cracks
filled with slickensided clay near its undulating top.

Interpretation: During the L, M, N, and 0 cycles, lacustrine sedimentation from suspension was interrupted by influxes of sand from
Gap Creek fan delta (site 9), probably in the form of low-density turbidity currents, reflecting high stream-discharge events. Wedging of
Ns sands and the absence of Ls and Ms sands from the transect (Fig.
14) suggest that sand transport was of variable distance or width.
The transient north-trending channels reflect sand supply from
Maple Creek. Temporary emersion of this part of the former lake
bed is indicated by slickensided cracks at the top of the Kcx bed. The
absence of the Ocx bed is attributed to erosion during and subsequent to lake drainage, which was followed by deposition of
sandy-loam loess.
Significance: During the early phase of Junction Lake (before
6800 BP), desiccation of the Kcx bed (approximately 755 m a.s.1.)
indicates the presence of a closed basin with lake level approximately 4 m below the mid-Holocene outflow level (between 759 and
760 m a.s.l., sites 12,13, and 16). Subsequent continuous lacustrine
sedimentation occurred until the close of the 0-cycle, with thick clay
beds probably reflecting prolonged lake stands below outlet level,
when there was no transmission of suspended sediment towards
Hanson Lake. These beds may therefore mark prolonged droughts.
Aggradation to about 757 m a.s.1. at this site is traceable to approximately 758 m a.s.1. at site 10, which is also the elevation of a
post-Mazama outlet channel at site 14. A rapid decline of the lake's
water storage capacity likely caused accelerated cutting of the Eagle
outlet system, which triggered the demise of the lake.

North-trending channels, scoured into the Lcx and Ncx beds, have
fills of alternating thin and medium rippled sand and massive clay
(Fig. A-12A). Flutecasts at the sandy base of the uppermost channel
are overlain by planar clay (Fig. A-12B).

Figure A-1 1.
Distribution of Junction Lake sediments along shore blufat site 11. Units explained in text.

Figure A-12. Fill of channel incised into Ncx clay bed at site 11. A) exposure between 25 and 40 rn; scale is
1.0 rn; GSC 1999-043H; B) flutecasts at base of channelfill; scale is 0.8 rn; GSC 1999-0431.Photographs
by W.J. Vreeken.

W.J. Vreeken

SITE 12 - Duffee's site (Fig. 6, A-13)


Geomorphic sugace: lake plain of Junction surface

SITE 13 -Dean's sites (Fig. 6, A-14, A-15)


Geomorphic surface: lake plain of Junction surface

complex

complex

Latitude: 49'57'35"N
Longitude: 109'30'52" W
Sugace elevation: 759 m a.s.1.

Latitude: 49O57'30"N
Longitude: 109'30'17"W
Surface elevation: 760 m a.s.1.

Description: A 1 m high shoreline cliff along Junction Reservoir


exposes laminated silt-clay rhythmites (Fld), overlain by coarse
sand with scattered pebbles (Gms) and capped by a veneer of massive sand (Sme). Similar sand extends discontinuously across the
adjacent land surface. The top of the section lies approximately
0.5 m lower than a relict shoreline (approximately 759 m a.s.l.),
marked by surficial cobbles and pebbles, about 10 m away on the
adiacent surface. Another relict shoreline. found 0.5 km to the west
at~pproximately765 m a.s.l., is marked by a cobble-strewn surface
that slopes away from notches in adjacent hillsides.

Description: Reservoir bluff sections (Fig A-14) reveal dianuct


(Drnm) overlain by crossA
6
and planar-laminated sand
units (Sr and Sh), and massive sand with rare cobble-size dropstones (Srnd)
and a very thin clay bed
(Fld). Locally, thin gravel
Frne
(Gm) directly overlies the
diamict. The sections are
capped by massive sandy
loam to loamy sand (Fme).
In section B, this upper unit
includes buried, severely
..
bioturbated, generally dark
. . Smd
.
soils and one bioturbated
tephra bed (Fig. A-15).

Interpretation: The 765 m a.s.1, paleoshore marks the final and


lowest limit of glacial Lake Maple Creek, and correlates with the
earliest and highest limit of postglacial Junction Lake. The 759 m
a.s.1, shoreline marks a younger limit of Junction Lake. The
rhythmites (Fld) record sedimentation in glacial Lake Maple Creek,
whereas the overlying pebbly sand (Gms)
reflects a littoral zone associated with the
759 m a.s.1. shore. The pebbles and cobbles
Likely represent current and wave reworking of
glacial diamict exposed nearby. The sand
veneer (Sme) is probably eolian and related to
nearby blowouts, but here has been reworked
Q
by wave action.

't

*759

a.s.'.

Figure A-13.
stratigraphic
section, site 12.

Significance: The site estimates a highstand of


Junction Lake at 759 m a.s.l., attained shortly
before its demise. This elevation is consistent
with the 759-760 m a.s.1. range f o r
mid-Holocene high lake levels inferred from
sites 13 and 16.

+
I

I!.'./

Interpretation:
Till
(Dmm) is erosionally overlain by lag gravel (Gm) and
glaciolacustrine deposits
(Sr, Sh, Smd, Fld) laiddown
in glacial L a k e M a p l e
Creek. Cyclic accumulation
of loess (Fme) occurred
both before and after deposition of the tephra, likely
Mazama (see site 15).

2-

,
2e <

o
5-

5-

-2cI
C

Dmm

O='==

t760 rn a.s.1.

3-

- a

*759 rn a s.1.

Figure A-14.
Stratigraphic section, site 13.

Significance: The uppermost glaciolacustrine sediments lie between 759 and 760 m a.s.l., below the lowest limit of
glacial Lake Maple Creek (765 m a.s.1.). The sequence may have
been eroded during the later phases of Junction Lake, as the elevation is consistent with the 759-760 m a.s.1. range for mid-Holocene
high lake levels inferred at sites 12 and 16.

Figure A-15. Buried cumulic soil A horizon with Scoyenia


ichnofacies, site 13. Pale burrow infillings consist ofMazama
tephra. Photograph by W.J. Vreeken. GSC 1999-0435

GSC Bulletin 534

SITE 14 - reservoir-dam channel site (Fig. 7, A-16)


Geomorphic surjkce: Eagle outlet of Junction surface

SITE 15 - Eagle camping site (Fig. 7, A-17)


Geomorphic suface: Eagle outlet of Junction surface

complex

complex

Latitude: 49O58' 11"N


Longitude: 10930'55" W
Suface elevation: 759 m a.s.1.

Latitude: 49O58'11"N
Longitude: 109'3 1'26"W
Suface elevation: 756 m a.s.1.

Description: Cliff exposure along Junction


Reservoir lies parallel to, and 2-3 m away
from, an approximately 30 m long dry channel with a base at approximately 758 m a.s.1.
The channel is oriented towards the former
Eagle outlet of Junction Lake. The section
reveals a boulder bed (Gm) draped by thin
massive sand (Smd) similar to site 13B. This
is successively overlain by gritty silt with
abundant tephra grains (T), thin massive
loamy sand ( F m e ) , and horizontal,
thin-bedded, sandy loam with pebble stringers that include small bones (Sh). The upper
loam forms the wall of the dry channel.

Description: Cliff on eastern valley wall of the Eagle spillway


(approximately 765 m a.s.1.) of glacial Lake Maple Creek exposes
basal loam diamict ( D m ) , overlain by gravel (Gm) and apparently
massive sand (Sme), and capped by massive sandy loam (Fme) with
buried soils. Biochannels from one buried soil are filled with tephra
(T), identified as Mazama by Westgate and Naeser (1985, site
UA-311).

Interpretation: Diamict (Drnm), exposed


nearby, is erosionally overlain by the boulder
bed (Gm) and by sand deposited in glacial
Lake Maple Creek, as at site 13B.The base of
the reworked tephra bed (T), likely Mazama
(see site 15), is probably erosional and the
loamy sand (Fme) represents sandy loess.
The bedded upper loam (Sh) is a channel
deposit.

The valley bottom is occupied by the Weir (747 m a.s.1.) and Bull
(745 m a.s.1.) fluvial surfaces, above which are several erosional surfaces, discernible along the western valley wall. The spillway limit is
marked by abundant surface cobbles and boulders.

Interpretation: The gravel bed (Gm) on till


( D m ) and nearby surface boulders are an
erosional lag formed by downcutting of the
glacial spillway. The spillway also served as
the Eagle outlet of postglacial Junction Lake
and evenlually evolved into Maple Creek
valley. The Sme and Fme units are likely
eolian, with accumulation beginning well
before 6800 BP and continuing thereafter.

t758 rn a.s.1.

Figure A-16.

section, site 14.

a;F

5
a
0

Significance: The channel (758-759 m a.s.1.) postdates 6800 BP but


predates the adjacent Weir tenace (748 m a.s.l.), now inundated by
the reservoir but recognizable on an old map (Prairie Farm Rehabilitation Administration, unpub, map, scale 1:9150, 1938). Its elevation lies slightly below the 759-760 m range for mid-Holocene high
levels of Junction Lake, inferred at sites 12, 13, and 16. The channel
base is about even with the top of the Ocx bed at site 10 (see also site
11). This channel probably formed part of the last outlet system of
Junction Lake, because any lower channel would have drained the
lake and been subsequently incised by Maple Creek.

Significance: All terrain to the north of this


site, extending to the Fox Valley end
moraine, is less than 765 m a.s.1. Thus, glacial meltwater from the north must have
reached glacial Lake Maple Creek via
, srne
ice-walled, probably subglacial, conduits.
2
Gm
Following site deglaciation, water flow was
Drnrn
reversed when the spillway became a
+ 756 rn a s I. postglacial lake outlet, subject to retrograde
incision. Because the tephra has been identified as Mazarna, this site is a reference for
Figure A-17tephra beds from similar sequences nearby.
O

Stratigraphic
section, site 15.

W.J. Vreeken

SITE 16 - Lightfoot outlet (Fig. 7)


Geomorphic surface: Lightfoot outlet of Junction surface
complex

Latitude: 49'58'02"N
Longitude: 109"30'02"W
Su@ce elevation: 760 m a.s.1.
Description: This hand-augered site lies on a saddle of a longitudinal depression (approximately 760 m a.s.1.) within the Lightfoot
spillway (765 m a.s.1.) of glacial Lake Maple Creek. The depression
marks the Lightfoot outlet of postglacial Junction Lake, and is connected to Maple Creek valley, towards the northwest, via a hanging
side-valley. At thatjunction, thls valley is approximately 2 m higher
than the Weir fluvial surface. At the site, gravel, at approximately
1.2 m depth, is successively overlain by thin-bedded fine sand, silty
clay loam, silt with tephra grains, silt loam, loamy sand, silt loam,
and sandy loam.
Interpretation: A basal gravel lag deposit is overlain by
thin-bedded fluvial deposits. The tephra grains are attributed to the
6800 BP Mazama eruption (see site 15).
Significance: The Lightfoot spillway evolved into an outlet of Junction Lake, similar to the history inferred for the Eagle spillway (site
15). The presence of waterlain deposits with tephra grains indicates
that the Lightfoot outlet remained in use after the Mazama eruption
(6800 BP). The elevation range for the channel fill (759-760 m
a.s.1.) is the same as that of mid-Holocene high levels of Junction
Lake (sites 12 and 13). This outlet was abandoned when a slightly
lower channel was formed at site 14. Abandonment of the Lightfoot
outlet is attributed to faster headward incision of the Eagle outlet after
6800 BP.

SITE 17 - Hanson surface sites (Fig. 7)


Geomorphic surface: lake plain of Hanson surface
Latitude: 5001'05"N
Longitude: 109'3 1'00" W
Surface elevations: 73 1 m a.s.1.
Description: The Weir alluvial surface grades into the Hanson lake
plain. The Bull surface lies 1-2 m below these surfaces, and includes
the modern creek channel. On the paleoshore of Hanson Lake plain
(site 17A, 732 m a.s.1.) and on topographic highs surrounded by the
lake plain, gravel and coarse sand occur at the surface. Within the
lake plain (site 17B, 731 m a.s.l.), along a manmade channel, basal
clast-supported gravel and coarse sand is successively overlain by
massive clay (Fc, 0.2 m), rippled fine sand (Sr, 1.6 m) with rare
shells from snails and bivalves, and massive clay (Fc, 0.4 m).

SITE 18 - Feil's crossing transect (Fig. 8, A-18)


Geomorphic surface: Bull surface
Latitude: 50'1 1'1 1"N
Longitude: 109O27'30" 5W
Surface elevation: 715 m a.s.1.
Description: Transect data are from three boreholes (sites 18A, B,
and C) on the Bull surface floodplain (715 m a.s.l.), and from an
abandoned borrow pit on the adjacent Neitz surface (18D; 716 m
a.s.1.).In the latter, sand with minor gravel is devoid of fine overburden. At sites 18A and C, basal gravel (Gm), draped by apparently
massive coarse sand, is overlain by massive gritty clay (Fcgm) and
clay (Fc), interbedded and interlaminated with very fine sand (Sh
and FI) with organic-enriched zones in the upper 1.5 m. At site 18B,
approximately 4 m of smooth clay (Fc) is overlain by 4.3 m of interbedded deposits similar to those at site 18A. An organic-enriched
zone at 5.9 m depth dates to 9500 BP (TO-4413).
Interpretation: Glaciofluvial sand and gravel underlies Neitz surface and a buried channel beneath the modern floodplain. Gritty clay
(Fcgm) at the base of the fill is a mixture of postglacial clay, originating upvalley, and reworked local material. The thick basal clay (Fc)
at site 18B is lacustrine and the interbedded upper complex reflects
lacustrine and fluvial sedimentation alternating with emersive
phases.
Significance: Aside from the Fcgm facies, the infilling of the buried
channel is devoid of sediment coarser than medium sand, and hence
texturally distinct from its substrate. This suggests that the channel
was cut by deglacial meltwater flowing from Bigstick basin (see also
site 19). Early Holocene sedimentation occurred in a lake that
extended ~ e l l ; ~ v a l lfrom
e ~ Feil's Crossing. Subsequent accumulation of alternating lacustrine and fluvial sediments reflects fluctuating lake levels, although the absence of a fine-grained overburden on
Neitz surface shows that postglacial lake and stream levels did not
exceed 716 m a.s.1. Because the water and sediment storage capacity
of the depository decreased throughout postglacial time, excess
inflow of water must have drained laterally into the glaciofluvial
substrate of Neitz surface.

Fcgm
Gm

Interpretation: Glaciofluvial sand and gravel formed the shores,


islands, and substrate of former Hanson Lake. At site 17B, this substrate is overlain by postglacial lacustrine clay interbedded with fluvial sand.
Significance: Neither Hanson Lake nor any postglacial stream in
this part of lower Maple Creek valley stood higher than approximately 732 m a.s.1. The north-trending valley postdates higher adjacent
east-trending ridges of glaciofluvial sediment (Fig. 3). Since evidence of postglacial stream activity, other than that associated with
the Weir and Hanson surfaces, is absent, the valley form likely
relates to a late deglacial phase of glaciofluvial erosion. Postglacial
streamflow, initially regulated by outflow from Junction Lake until
its demise, after 6800 BP, caused aggradation of the Hanson Lake
plain. The demise of Hanson Lake terminated the Weir stream cycle
(ca. 1500 BP; see site 8).

Figure A-18.
Stratigraphic section, site 18.

GSC Bulletin 534

SITE 19 - Bigstick inlets area (Fig. 8, A-19)


Geomorphic surface: lake plain of Bigstick surface
Latitude: 5012'N
Longitude: 109O25'W
Surface elevation: 715 m a.s.1.
Description: 'The lake plain (approximately 715 m a.s.1.) features
dry shallow channels traversing clay flats, and lies below the adjacent Neitz surface (716 m a.s.1.). The core from site 19A, a dry channel bottom, consists of coarse sand on pebble gravel (Om),overlain
by 1.5 rn of thin clay (Fc) and interlaminated clay and fine sand (FI).
Cores from sites 19B, C, and D illustrate subsurface topographic
variability, with at least 4.5 m of relief on the sand and gravel (Gm)
substrate. The overburden is thin gritty clay (Fcgm) and thick
smooth clay (FC;4.6 rn at site 19D), aid is-capded at s h 1 9 ~
interbedded and interlaminated with silt and fine sand (FI), similar to
the thin cover on basal sand at site l9B. Core 19E, from the western
channel, consists of thick smooth clay (Fc) overlain by approximately 2 m of interbedded deposits (F1). Approximately 250 m east of
site 19A, a dugout on Neitz surface reveals coarse sand overlain by
0.5-1.0 m of massive sandy loam. A scour depression along the edge
coarse sand
of ~ e i t surface,
z
and approximately 1 m lower,
overlain by 0.4 rn of fine sandy clay and clay.

Amorphous organic material from the top of the thick clay at site
19E was dated at 7500 BP (TO-4410). Two apparently anomalous
dates of 4700 BP (TO-4411) and 6200 BP (TO-4411R) were
obtained from a basal organic-enriched zone at the same site.
Interpretation: Glaciofluvial sand and gravel underlying Neitz surface is directly overlain by discontinuois sandy loess. 1;also forms
the substrate of fine-grained deposits beneath the Bigstick Lake
plain. Postglacial deposits consist of interbedded lacustrine clay and
fluvial sand-clay complexes. The absence of coarse sand in
postglacial deposits indicates that relief on the basal sand-gravel
substrate must be attributed to erosion by meltwater discharged from
Bigstick basin. The anomalous basal dates for site 19E likely reflect
contamination during coring.
of the lake plain
significance: l-he buried glaciofluvial
processes. Postglacial lacustrine sedirecords deglacial
mentation ofclay occurred in a vast lake, beginning prior to 9500 BP
(site 18B) and lasting until ca. 7500 BP (site 19E). This indicates
that, until ca, 7500 BP, Bigstick basin received ample streamflow,
suggesting that neither Junction nor Hanson lake had prolonged low
water levels during this interval. Subsequent sedimentation
involved alternating fluvial and lacustrine deposition, with ~ e c u m n t
high lake levels. Yet, the highest postglacial stand of Bigstick Lake
was less than 716 m a.s.1. ~sinterpretedat site 18, inflow exceeding
the diminishing storage capacity of the lake must have seeped into
surrounding coarse-grained deposits.

Figure A-19.
Stratigraphic section, site 19.

AUTHOR INDEX
Aitken, A.E. . . . . . . . . . . . . . . . . . . . . . . 173
(email: aaitken@arts.usask.ca)
Basinger, J.F. . . . . . . . . . . . . . . . . . . . . . 139
(email: jim.basinger@usask.ca)
Birks, S.J. . . . . . . . . . . . . . . . . . . . . . . 57,81
(email: sjbirks@sciborg, uwaterloo.ca)
Burt, A.K. . . . . . . . . . . . . . . . . . . . . . . . 173
(email: akburt@scimail.uwaterloo.ca)
David, P.P.. . . . . . . . . . . . . . . . . . . . . . . 223
(email: david@ere.umontreal.ca)
Goldsborough, L.G.. . . . . . . . . . . . . . . . . . 111
(email: ggoldsb@umanitoba.ca)
Hal1,R.I.. . . . . . . . . . . . . . . . . . . . . . . . 125
(email: rihall@scimail.uwaterloo.ca)
Huntley, D.J. . . . . . . . . . . . . . . . . . . . 21 1,223
(email: huntley @sfu.ca)
Last, M. . . . . . . . . . . . . . . . . . . . . . . . . .23
(email: mlast@ms.umanitoba.ca)
Last, W.M. . . . . . . . . . . . . . . . . . . 95,125,173
(email: mlast@ms.umanitoba.ca)
Leavitt, P.R. . . . . . . . . . . . . . . . . . . . . . . 125
(email: leavitt@uregina.ca)
Lemmen, D.S.. . . . . . . . . . . . . . . . 1,7,199,223
(email: dlemmen@NRCan.gc.ca)
Lian, O.B. . . . . . . . . . . . . . . . . . . . . . . . 211
(email: Olav.Lian@vuw.ac.nz)

Nelson, H.L. . . . . . . . . . . . . . . . . . . . . . . 257


(email: hugh.nelson@ec.gc.ca)
Remenda, V.H. . . . . . . . . . . . . . . . . . . . 57,81
(email: remenda@geol.queensu.ca)
Richmond, K.-A. . . . . . . . . . . . . . . . . . . . 111
(email: karich@ibm.net)
Sauchyn, D.J. . . . . . . . . . . . . . . . . 239,249,257
(email: sauchyn@leroy.cc.uregina.ca)
Shang, Y. . . . . . . . . . . . . . . . . . . . . . . . . 95
(email: shang@cc.umanitoba.ca)
Smol, J.P. . . . . . . . . . . . . . . . . . . . . . 67,125
(email: smolj @ biology.queensu.ca)
Spence,C.D . . . . . . . . . . . . . . . . . . . . . . . 249
(email: chris.spence@ec.gc.ca)
Vance,R.E.. . . . . . . . . . . . . . . . . . . . 1,7,125
(email: rvance@NRCan.gc.ca)
Vinebrooke,R.D. . . . . . . . . . . . . . . . . . . . 125
(email: vinebrro@meena.cc.uregina.ca)
Vreeken, W.J. . . . . . . . . . . . . . . . . . . . . . 267
(email: vreekenw@qsilver.queensu.ca)
Wilson, S.E. . . . . . . . . . . . . . . . . . . . . 67,125
(email: wilsons@biology.queensu.ca)
Wolfe, S.A. . . . . . . . . . . . . . . . . . . . . 199,223
(email: swolfe@NRCan.gc.ca)
Yansa,C.H . . . . . . . . . . . . . . . . . . . . . . . 139
(email: chyansa@students.wisc.edu)

Вам также может понравиться