Вы находитесь на странице: 1из 32

The colour of climate: changes in peat

decomposition as a proxy for climate change


a study of raised bogs in south-central Sweden

Anders Borgmark

Avhandling i Kvartrgeologi
Thesis in Quaternary Geology
No. 4
Department of Physical Geography and Quaternary Geology
Stockholm University
2005

Cover artwork by Jens Heimdahl


Decomposing Sphagnum leaf

ISBN 91-7155-101-8
ISSN 1651-3940
Akademitryck AB

The colour of climate: changes in peat


decomposition as a proxy for climate change
a study of raised bogs in south-central Sweden

Anders Borgmark

This doctoral thesis consists of four papers and a synthesis. The four papers
are listed below and are referred to as Paper I-IV in the text:
Paper I:
Gunnarson, B. E., Borgmark, A. and Wastegrd, S. 2003: Holocene
humidity uctuations in Sweden inferred from dendrochronology and peat
stratigraphy. Boreas 32, 347-360.
Reprinted with permission by Taylor & Francis AS.
Paper II:
Borgmark, A. 2005: Holocene climate variability and periodicities in
south-central Sweden, as interpreted from peat humication analysis.
The Holocene 15, 387-395.
Reprinted with permission by Arnold Journals.
Paper III:
Borgmark, A. and Schoning, K. in press: A comparative study of peat proxies
from two eastern central Swedish bogs and their relation to meteorological
data. Journal of Quaternary Science.
Copyright 2005, John Wiley & Sons, Ltd.
Paper IV:
Borgmark, A. and Wastegrd S.: Regional and local patterns of peat
humication in three raised peat bogs in Vrmland, south-central Sweden.
Manuscript.

Abstract
This thesis focuses on responses in raised bogs to changes in the effective humidity during the Holocene.
Raised bogs are terrestrial deposits that can provide contiguous records of past climate changes. Information
on and knowledge about past changes in climate is crucial for our understanding of natural climate variability.
Analyses on different spatial and temporal scales have been conducted on a number of raised bogs in southcentral Sweden in order to gain more knowledge about Holocene climate variability.
Peatlands are useful as palaeoenvironmental archives because they develop over the course of millennia and
provide a multi-faceted contiguous outlook on the past. Peat humication, a proxy for bog surface wetness,
has been used to reconstruct palaeoclimate. In addition measurements of carbon and nitrogen on sub-recent
peat from two bogs have been performed. The chronologies have been constrained by AMS radiocarbon dates
and tephrochronology and by SCPs for the sub-recent peat.
A comparison between a peat humication record from Vrmland, south-central Sweden, and a dendrochronological record from Jmtland, north-central Sweden, indicates several synchronous changes between
drier and wetter climate. This implies that changes in hydrology operate on a regional scale.
In a high resolution study of two bogs in Uppland, south-central Sweden, C, N and peat humication have
been compared to bog water tables inferred from testate amoebae and with meteorological data covering the
last 150 years. The results indicate that peat can be subjected to secondary decomposition, resulting in an apparent lead in peat humication and C/N compared to biological proxies and meteorological data.
Several periods of wetter conditions are indicated from the analysis of ve peat sequences from three bogs
in Vrmland. Wetter conditions around especially c. 4500, 3500, 2800 and 1700-1000 cal yr BP can be correlated to several other climate records across the North Atlantic region and Scandinavia, indicating wetter
and/or cooler climatic conditions at these times. Frequency analyses of two bogs indicate periodicities between 200 and 400 years that may be caused by cycles in solar activity.
Keywords: Holocene, south-central Sweden, peat humication, bog hydrology, climate variation, spectral
analysis
Anders Borgmark, Department of Physical Geography and Quaternary Geology.
Stockholm University, SE-10691 Stockholm, Sweden
E-mail: anders.borgmark@geo.su.se

The colour of climate: changes in peat decomposition as a proxy for climate change

Introduction
The geographic position of Scandinavia with the
North Atlantic Ocean to the west and the Eurasian
continent to the east, with its cold air masses, causes
high climate variability and large and often rapid
weather changes (Alexandersson and Andersson,
1995; Vedin, 2004). The proximity to the ocean is
a major factor affecting climate (Bradley, 1999) and
the North Atlantic climate system with strong westerly winds and warm sea water controls much of the

Scandinavian weather and climate. Understanding


the systems controlling climate conditions around
the North Atlantic is central in climatic research
(Marshall et al., 2001).
In order to improve the understanding of climate
variability and its underlying causes, reconstructions
of palaeoclimate are fundamental (e.g., Alverson
et al., 2001; Bradley et al., 2003a; Oldeld, 2004;
Moberg et al., 2005). Variations in climate are recorded in natural archives such as ice sheets and

Figure 1. Map of Scandinavia with investigated and analysed bogs in this study. Studied sites are
marked by lled circles, open circles represent sites that have been excluded; M: Mosstakanmossen,
K: Klaxsjmossen and B: Brrudsmossen. The position of Lake Hckren, the study area for dendrochronology (Paper I) is shown as a square. Approximate location of provinces mentioned in the text
is shown; Vrmland (V), Uppland (U), Smland (S), stergtland () and Jmtland (J).

Anders Borgmark

Figure 2. The Blytt-Sernander Scheme of postglacial climate periods in


continental north-western
Europe, 14C ages from
Mangerud et al. (1974).

glaciers, tree rings, deep sea- and lake sediments,


corals, speleothems and peatlands (e.g., Barber,
1982; Berglund, 1986; Lowe and Walker, 1997; Bianchi and McCave, 1999; Cronin, 1999; Holmgren
et al., 1999; Lauritzen and Lundberg, 1999b; Briffa,
2000). These archives contain proxies of different
types; biological, physical, chemical and lithological
(Lowe and Walker, 1997) and analyses of these make
it possible for the scientic community to interpret
past climate changes and contribute to the understanding of Earths climate system. Although peatlands have long been used as records of past climate
change, research on peatlands still receives relatively
little attention within the palaeoclimatology community (e.g., Lowe and Walker, 1997; Bradley, 1999;
Cronin, 1999; Bradley et al., 2003b). It is noteworthy, however, that indices of considerable climatic
changes during the Holocene were rst registered
in peat stratigraphical records (e.g., Blytt, 1876;
Sernander, 1908; Weber, 1926). It was not possible to
discern such changes in the early ice and marine core
records and were therefore not acknowledged until
recently when high resolution records (e.g., OBrien
et al., 1995; Bond et al., 1997) conrmed the peat
based indices (Chambers and Charman, 2004).
This thesis focuses on Holocene climatic changes
in Sweden and how such changes are recorded in
peat sequences of raised bogs (Figure 1). The main
objectives are to:
Study changes in peat decomposition to infer
past changes in bog hydrology on different spatial
and temporal scales.
Compare and correlate changes in bog hydrology
with other proxies of climate change in order to evaluate on which scale, both spatial and temporal, the
peat-proxies from south-central Swedish bogs could
be a useful tool for palaeoclimatic reconstructions.
Assess the causes of climate change leading to
changes in bog surface wetness and the extent to
which precipitation and temperature control peat
decomposition processes.
In this thesis palaeohydrological changes in bogs
have been determined primarily through measure-

ment of the degree of peat decomposition. As a supplement to this method, measurements of organic
carbon and nitrogen have been used together with
analyses of testate amoebae assemblages on two
sub-recent peat sequences. The chronologies of the
analysed peat sequences have been established by a
combination of mainly AMS-radiocarbon dates and
tephrochronology.
Peat as a palaeoclimatic archive
Peat has been used as an archive for climate variations for over a hundred years. Scandinavian peatlands provided the basis for the rst Holocene climatostratigraphy (Chambers and Charman, 2004).
The climate periods of the Holocene (Blytt, 1876;
Sernander, 1908), which have become known as the
Blytt-Sernander scheme (Figure 2) are still included
in the Scandinavian chronostratigraphy (Mangerud
et al., 1974) and less formally also in use in other
parts of Europe (Chambers and Charman, 2004).
The pioneer work of e.g., Blytt (1876), Sernander
(1908, 1909, 1912) and Weber (1926) was followed
by several investigations in Sweden (e.g., von Post
and Sernander, 1910; von Post and Granlund, 1926;
Granlund, 1932), which established peatland stratigraphy as a tool for the interpretation of past climate
changes.
At the Geological Congress in Stockholm in 1910
the hypothesis of a cyclic regeneration of peat growth
i.e., from hollow to hummock and back to hollow,
was presented (von Post and Sernander, 1910). This
and other hypothesises, such as the one on autogenic
succession of peat (Osvald, 1923), led to a misconception on how ombrotrophic peatlands develop
(Backus, 1990). For many years the predominating
view was that peatlands grow primarily through autogenic processes and not under the control of allogenic processes (i.e., climate). The focus on autogenic
processes led to a decline in peat-based palaeoclimatic science (Chambers and Charman, 2004).
The view of cyclic regeneration as the primary
process for peat accumulation in ombrotrophic
mires was challenged in the early 1980s (Barber,
1981; Frenzel, 1983). Barber (1981) rejected the
hypothesis by extensive plant macrofossil analyses
of peat stratigraphies along transects, showing that
the hummock/hollow complexes could be stationary for long periods of time, and he concluded that
the growth of a bog is to a large extent controlled
by climate. Barber (1981) concluded that different
thresholds could inuence regional as well as local
variations in bog growth, but factors such as hydrology, drainage, plants life cycles and pool-size are all

The colour of climate: changes in peat decomposition as a proxy for climate change

Table 1. Compilation of Granlunds (1932) recurrence


surfaces, with their relative occurrence throughout his
investigation area (1 least number of surfaces found and
5 most number found). Recurrence surfaces VI, VII were
described by Sandegren and Magnusson (1937).

Recurrence
surface
I
II
III
IV
V
VI*
VII*

Age
AD/BC
AD 1200
AD 400
600 BC
1200 BC
2300 BC
2900 BC
3700 BC

Cal yr
BP
750
1550
2550
3150
4250
4850
5650

Relative
occurrence
1
2
5
3
4
N/A
N/A

subordinate to climate.
Today a wide range of methods are used to reconstruct past climate from peat stratigraphies (Chambers and Charman, 2004). Quantitative macro- and
microfossil analysis (e.g., Charman and Warner,
1997; Barber et al., 1998, 1999; Charman et al.,
2000; Hughes et al., 2000; Mauquoy et al., 2004a)
together with peat humication (e.g., Nilssen and
Vorren, 1991; Blackford and Chambers, 1993;
Caseldine et al., 2000; Langdon and Barber, 2004;
Roos-Barraclough et al., 2004; Blundell and Barber,
2005) are the most widely used methods. The range
of biological, physical and chemical methods that
are now available make multiproxy investigations
possible. Although some of the proxies may not be
independent of each other e.g., a plant-species signal
within the degree of humication, they can nevertheless be used as high resolution proxies with excellent possibilities for interpretation (Chambers and
Charman, 2004).
As mentioned above, most of the early peat studies
were performed in Scandinavia. Blytt (1876) and
Sernander (1908, 1909, 1912; von Post and Sernander, 1910) conducted pioneer research, which was
followed by von Post and Granlund with the inventory of peatlands in southern Sweden (von Post, 1921;
von Post and Granlund, 1926; Granlund, 1932).
These extensive spatial and stratigraphical studies
led to Granlunds thesis (1932) on the geology of
raised bogs in Sweden. In his thesis he studied the relationship between the dome-shape of bogs and the
amount of precipitation in different areas of southern Sweden, investigated the capillarity of different
types of Sphagnum peat and formulated the concept
of recurrence surfaces (Swedish: rekurrensyta, RY).
A recurrence surface is a distinct change from dark
well humied peat, representing drier conditions on
the bog, to light less humied peat, representing wetter conditions (Granlund, 1932). Granlund identi-

Figure 3. Southern Sweden with relative occurrence of


bogs with recognised recurrence surfaces (RY) during the
peat inventory of southern Sweden (1917-1923), compiled from data in Granlund (1932). White represents
areas with occasional bogs with recognised RYs. Light
grey denotes moderate occurrence of bogs with RYs and
dark grey represent areas with an high occurrence of bogs
with recognised RYs. The peat inventory was restricted to
south and south-central Sweden with the exception of the
islands land and Gotland in the Baltic Sea. The northern
limit of the inventory is marked by a thick line.

ed ve different surfaces in southern Sweden (Table


1), but the number varied throughout the investigated area (Figure 3). Most recurrence surfaces were
found in Smland and further northwest towards
Vrmland (Granlund, 1932). Two older recurrence
surfaces were theoretically deduced by G. Lundqvist
(1932) and later described by e.g., Sandegren and
Magnusson (1937).
Detailed stratigraphical work on peatlands continued during the regional mapping of Quaternary
deposits (e.g., J. Lundqvist, 1958a). During the late
1940s and 1950s the radiocarbon dating technique
was developed and tested on peat. However, this
novel dating method also introduced a problem: the
radiocarbon ages did not exactly t into the previously established chronology (e.g., J. Lundqvist,
1957; Mller and Stlhs, 1964). Subsequently,
the ages of known recurrence surfaces were interpreted as being asynchronous (J. Lundqvist, 1957).
Therefore, the concept of recurrence surfaces as indicators of regional climate changes was questioned

Anders Borgmark

Bog
Oxic

Boun

ic
Anox

su rf ace

ACROTELM
ne
dary zo

High de
comp
.
Fluctuating
water table

CATOTELM

Low dec
omp.

Figure 4. Schematic picture showing the position of the


acrotelm, where the active decomposition takes place, and
the catotelm, with very a low decay rate, in relation to the
uctuating groundwater table near the bog surface. The
uctuations of the groundwater table determine the depth
of the acrotelm.

(G. Lundqvist, 1962). The scientic community in


Finland has been sceptical to the notion of climatic
inuence on the development of recurrence surfaces
since the theory was introduced (e.g., Aario, 1932;
Tolonen et al., 1985; Tolonen, 1987) and is, according to Korhola (1992), still placing major emphasis
on autogenous succession. Due to the national and
international scepticism, recurrence surfaces became
considered to be of minor use as indicators of palaeohydrological changes and as exact stratigraphical
markers in Sweden (Fries, 1951; J. Lundqvist, 1957;
G. Lundqvist, 1962; Nilsson, 1964). Despite the
general decline in peat-based studies a number of
investigations have been conducted in Fennoscandia regarding different aspects of peat growth, peat
accumulation rates, bog development, local environment and degree of humication (e.g., Nilsson,
1964; Tolonen, 1973, 1987; Aaby and Tauber, 1974;
Aaby, 1986; Foster et al., 1988; Svensson, 1988).
Aaby (1976) and Aaby and Tauber (1974) made attempts to identify cyclic variations in the climatic
signal from raised bogs in Denmark and this important work has been followed by a number peatbased studies on palaeoclimate in Scandinavia (e.g.,
Thelaus, 1989; Nilssen and Vorren, 1991; Korhola,
1992, 1995; Oldeld et al., 1997; Mauquoy et al.,
2002; Barber et al., 2004; Bjrck and Clemmensen,
2004; Bergman, 2005; Schoning et al., 2005). Bar-

ber et al. (2004) stated that there is a need for more


peat-based palaeoclimatic research from this part of
Europe and several bogs in Sweden seem to provide
valuable records for palaeoclimatic reconstructions
(Svensson, 1988; Bjrck and Clemmensen, 2004;
Bergman, 2005; Schoning et al., 2005).
Bog hydrology and decay
Peatlands are useful as palaeoenvironmental archives
because they develop over the course of millennia and
provide a multi-faceted continuous outlook on the
past. In order to accumulate peat an excess of water
is necessary to prevent complete decay of the biomass produced by the growth of plants (Charman,
2002). Because of the importance of water in peatlands their hydrology has become an important area
of study, and early studies on this subject were conducted in Russia by e.g., Romanov (1968), whose
ideas have been incorporated in the present standard
work on peatland hydrology (Ingram, 1983). Ingram
set up what is known as the groundwater mound
hypothesis (Ingram, 1982, 1983) which seeks
to explain the size and shape of raised mires. Until recently it was thought that capillary action was
responsible for holding the water table above the
surrounding landscape and that the increasing peat
column moves the groundwater table up through the
bog (Charman, 2002). However, this hypothesis was
considered already in the 1930s when Granlund
(1932) conducted an investigation on the capillarity
of low humied Sphagnum peat and concluded that
the capillary rise of water was insufcient to create
a groundwater mound. Today the domed prole is
thought to be mainly a result of low hydraulic conductivity in the catotelm i.e., the lower anoxic part
of the peat column (Ingram, 1982, 1983). In order
to improve models of peatland groundwater hydrology the rather static model of Ingram (1982, 1983)
has to be developed into more dynamic models, e.g.,
incorporating models of decay and different values
for conductivity (Charman, 2002).
The decay of plant material mainly takes place in
the upper aerated layer of the peat (Figure 4, Table 2)
i.e., the acrotelm (Ingram, 1978, 1983), being largely
the result of microbial decomposition and larger soil

Table 2. Main properties of the acrotelm and catotelm in peatlands. Modied from Charman
(2002).

Property
Water content and movement
Oxygen supply
Microbial activity
Decay rate

Acrotelm
Variable, rapid movement
Aerobic (periodically)
High
Rapid

Catotelm
Constant, very slow movement
Anaerobic
Low
Slow

The colour of climate: changes in peat decomposition as a proxy for climate change

animals (e.g., protozoa and nematodes) (Charman,


2002). The decay rate in the acrotelm varies considerably between sites and conditions, and the major
factors controlling the decay rates are water content,
temperature, oxygen supply, plant material and microbial and soil animal populations (Clymo, 1983;
Charman, 2002). In the catotelm (Figure 4, Table 2)
these decay processes are practically absent with a
decay rate estimated to be only c. 0.1 % of that in the
acrotelm (Ingram, 1978; Belyea and Clymo, 2001).
The roles and interactions of the different factors affecting the decay rate is complex and are not yet fully
understood (see Charman [2002] for a review). The
boundary between the acrotelm and catotelm is not
sharp but corresponds roughly to the lowest level of
the groundwater table in the summer (Clymo, 1984).
The aerated acrotelm is the contributor of organic
material to the catotelm and in the process of transferring material from the aerobic layer to the anaerobic layer about 80-95% of the total organic production is lost (Clymo, 1984; Warner et al., 1993).
Since plant decay is almost entirely conned to the
acrotelm, the thickness of this zone largely determines how decomposed the peat will become. The
thickness of the acrotelm is considered to be equal to
the degree of surface wetness and the decomposition
of peat is therefore primarily a function of the degree
of surface wetness (Caseldine et al., 2000). Surface
wetness is in turn mainly a function of precipitation
and evapotranspiration, but the precise relationship
is not known (Charman, 2002). There are several
factors, both autogenic and allogenic, affecting the
surface wetness and moisture of raised bogs at different scales (Figure 5), of which climate acts on the
Microtopography

Microclimate

Vegetation

Succesion

Mire expansion

Human impact

Climate
(precipitation evapotranspiration)

Regional groundwater

Figure 5. Model of factors affecting bog surface wetness at


different spatial scales. Microscale processes are principally affected by autogenic factors, at larger scales is surface
wetness more likely affected by allogenic factors. Redrawn
from Charman (2002). John Wiley and Sons Ltd, reproduced with permission.

broadest spatiotemporal scale implying that replicate changes across sites can normally be attributed
to changes in climate (Charman, 2002).
Holocene climatic changes
During the Holocene there have been a number of
rapid changes in climate in the Northern Hemisphere
(e.g., Mayewski et al., 2004; Snowball et al., 2004).
The early Holocene (c. 11,500-7500 cal yr BP) was
characterised by a quite variable climate; large ice
sheets still covered parts of the Northern Hemisphere and changes in ice sheet extent, mass balance
and large temperature differences between the warm
Atlantic waters and the cooler air masses circulating over the Northern Hemisphere caused this climatic instability (Mayewski et al., 2004; Snowball
et al., 2004). Despite this instability, a general warming trend occurred during the early Holocene. Sea
surface- and atmospheric temperatures may have
reached their Holocene maximum during the early
Holocene, indicated as e.g., c. 3C higher temperature at the summit of the Greenland ice sheet (compared to the last 500 yr) (Dahl-Jensen et al., 1998)
and c. 1.5-2.1C warmer than today in northern
Scandinavia (Barnekow, 2000; Larocque and Hall,
2004). However, three cooling events were superimposed on the general warming trend, of which
the 8200 yr event (Alley et al., 1997; Bond et al.,
1997) probably was the most striking (Snowball et
al., 2004).
The Holocene Thermal Maximum, which experienced c. 2C higher summer temperatures compared
to today in Fennoscandia, occurred between c. 7500
and 5000 cal yr BP (Snowball et al., 2004). Different
palaeoclimate records indicate that this was mainly
a dry period in Fennoscandia (e.g., Digerfeldt, 1988;
Eronen et al., 1999; Sepp et al., 2002). Towards the
end of the thermal maximum a cooling was commenced at c. 6000 cal yr BP (Mayewski et al., 2004);
this cooling is evident in e.g., the GISP2 record as
an increase in Na+ and sea salt-dust (OBrien et al.,
1995; Mayewski et al., 1997), ice rafted debris (IRD)
in the North Atlantic (Bond et al., 1997; Bond et al.,
2001) and glacier advances in Scandinavia (Karln
and Matthews, 1992; Karln and Kuylenstierna,
1996; Nesje et al., 2001) (Figure 6).
Around 4000 cal yr BP a number of wet-shifts
are observed in many raised bogs across northwestern Europe (Anderson, 1998; Anderson et al.,
1998; Hughes et al., 2000; Barber et al., 2004) (Figure 6) and several other records indicate a shift towards more variable conditions (Bond et al., 1997;
Barnekow, 2000; Mayewski et al., 2004; Snowball

Anders Borgmark

Figure 6. Examples of palaeoclimate series from the Northern Hemisphere for the last 10,000
calendar years. a) Timing of global Rapid Climate Change (RCC) (Mayewski et al., 2004) b)
200 yr smoothed GISP2 Na+ ion proxy for the Icelandic Low (Mayewski et al., 1997) c) Periods of increased levels of sea-salts and dust in the Greenland ice-cores (expansion of northern
polar vortex/increased meridional airow)(OBrien et al., 1995) d) IRD (Ice Rafted Debris)
events in North Atlantic marine sediments, indicative of cooling events (Bond et al., 1997) e)
Periods of glacier advances in Scandinavia (Denton and Karln, 1973) f) Periods of increased
number of wet-shifts found in peat bogs in western Europe (Hughes et al., 2000; Barber et
al., 2004) g) 200 yr smoothed 14C residuals (Stuiver et al., 1998) a global proxy for solar
variability. h) Composite record of peat humication from Vrmland, south-central Sweden
(average of humication index; Paper IV).

et al., 2004; Jessen et al., 2005). In Finland a period of more active peat initiation, and subsequent
spread, has been recognized 4300-3000 cal yr BP
(Korhola, 1992, 1995). This period of cooler and
possibly wetter conditions is followed by a period
of climate variability between 3500 and 2500 cal yr
BP (Figure 6). This period includes a number of wetshifts and periods of wet conditions recorded in peat

10

records (Hughes et al., 2000; Langdon et al., 2003;


Barber et al., 2004) and a generally colder climate
has been reconstructed in Finnish dendrochronological records (Eronen et al., 1999) and from chironomid assemblages in northern Sweden (Larocque and
Hall, 2004).
Between c. 1500 and 1000 cal yr BP many records
indicate a transition to cooler climate (e.g., Denton

The colour of climate: changes in peat decomposition as a proxy for climate change

and Karln, 1973; Karln and Matthews, 1992;


Bond et al., 1997; Mayewski et al., 1997, 2004;
Hughes et al., 2000; Langdon et al., 2003; Barber
et al., 2004)(Figure 6). During the Medieval Warm
Period (c. 1000-750 cal yr BP), however, temperatures could have been as much as 0.5-0.8C higher
than those today (Bond et al., 2001; Snowball et
al., 2004). Tree ring records indicate that summer
temperatures in high Northern Hemisphere latitudes were higher than the 20th century mean (Briffa,
2000) and European documentary records contain
evidence of mild winters during portions of this time
(Pster et al., 1998). There are also indications of
this being a period of lower precipitation which may
be of greater signicance than the temperature as an
indicator of a warm Medieval period (Bradley et
al., 2003b).
The last 500 years have been characterized by a
cooler climate (Mayewski et al., 2004) with advancing glaciers in Scandinavia and elsewhere (Karln
and Kuylenstierna, 1996) and increasing levels of
sea salt-dust in the Greenland ice sheet (OBrien et
al., 1995). However, there is some differences in the
timing and magnitude of regional variations in Scandinavia during the last 2000 years and there is only
a limited synthesis of high-latitude records available
(Bradley et al., 2003b; Snowball et al., 2004).
Changes in climate have many underlying causes,
of which many remain unknown and the effects on
climate are often poorly constrained. Both external
and internal forcing causes changes in an extremely
complex climatic system (Emeis and Dawson, 2003).
Lately solar variation has emerged as an important
external forcing mechanism for climatic variability
on decadal to millennial timescales (Chambers et
al., 1999; Bond et al., 2001) and changes in 14C
have been attributed to the changes in solar activity
(e.g., Karln and Kuylenstierna, 1996; Stuiver and
Braziunas, 1998; Chambers et al., 1999; van Geel
et al., 1999; Beer et al., 2002). A number of correlations have been made between climate proxy
records and solar variability in terms of 14C and
10
Be values (e.g., Denton and Karln, 1973; Reid,
1987; OBrien et al., 1995; Stuiver et al., 1998b; van
Geel et al., 1999, 2000; Bard et al., 2000; Bjrck
et al., 2001; Blaauw et al., 2004; Mauquoy et al.,
2004b; Mayewski et al., 2004). Exactly how the
relatively small changes in solar irradiance can affect
the climate is not clear, possible forcing mechanisms
have been discussed e.g., changes in ocean circulation (Bond et al., 2001), variations in UV irradiance
leading to altered production of ozone and absorption of heat in the atmosphere and as a consequence
shifts in the atmospheric circulation (van Geel et al.,

1999; Blaauw et al., 2004) and inuence of cosmic


rays on cloud formation (Carslaw et al., 2002). In
this study, peat-based palaeorecords are compared
to the reconstructed solar variation (Stuiver et al.,
1998b) in order to nd a possible cause for changes
in surface wetness on bogs (Figure 6g).
Geographical setting
The main study area is Vrmland (Figure 1) located
in south-central Sweden (Paper I, II and IV). The
region is heavily forested with mainly conifers (i.e.,
Scots pine and Norway spruce) and is intersected by
north-south fault-lines with long and often narrow
river valleys and elongated lake systems. The bedrock consists mainly of granites and gneisses and the
Quaternary deposits are dominated by till, glaciouvial deposits (eskers and deltas), numerous peatlands
and in lower areas clays (J. Lundqvist, 1958a). The
area was deglaciated between 10,100-9200 14C yr BP
(c. 11,000-9800 cal yr BP) and the highest shoreline
is situated between 180-200 m a.s.l. and increases
in elevation towards the north (J. Lundqvist, 1958a,
1994; Wastegrd, 1998).
In Vrmland a number of distinct shifts in peat humication recurrence surfaces have been reported
(Granlund, 1932; J. Lundqvist, 1957, 1958a, 1958b;
Persson, 1966) and the geographical/climatological
setting is thought to be favourable for palaeoclimatic investigations based on peat stratigraphy and the
occurrence of several tephra horizons, which make
detailed comparison between sites possible (Persson,
1966; Boygle, 1998; Zilln et al., 2002). In order to
facilitate selection of suitable peatlands the majority
of the sites were chosen from the extensive survey of
Quaternary deposits made by J. Lundqvist (1958a).
Since some of these bogs have since been drained,
cut or disturbed in other ways other bogs were also
cored and analysed. In paper I, II and IV humication data obtained from Stmyren, Kortlandamossen
and Fgelmossen have been used (Figure 1). Paper III
describes the bogs ltabergsmossen and Gullbergbymossen in Uppland, south-central Sweden (Figure 1), where the climate during the last 150 years
was investigated. In stergtland, sites described by
Granlund (1932) were initially chosen for study together with bogs chosen directly from the map of
Quaternary deposits (Svantesson, 1981). Many bogs
have, however, been disturbed by anthropogenic
activities in this region and only one sequence was
analysed. ngstugsmossen (N 58 50.8, E 14 16.5;
130 m a.s.l.; Figure 1) is a relatively small bog, about
500x300 m, with an undulating irregular border to
the surrounding forest. There are some indications

11

Anders Borgmark

15
10

60

5
40

20
0

-5
Jan Feb Mar Apr MayJun Jul AugSep Oct NovDec

100
80

100

5
40

20
0

-5
Jan Feb Mar Apr MayJun Jul AugSep Oct NovDec

-10

Precipitation (mm)

60

20

20

10

10
5

15

80

-5

Jan Feb Mar Apr MayJun Jul AugSep Oct NovDec

c) Herrberga

-10

20

T: +6.0C
P: 522 mm

15
10

60

5
40

20
0

-5
Jan Feb Mar Apr MayJun Jul AugSep Oct NovDec

of anthropogenic disturbance around the bog, but


no evidence of large drainage was observed. Two
peat sequences from the Faroe Islands were also
analysed but were excluded from the study due to a
problematic stratigraphy in one case and low temporal resolution for most of the Holocene in the other
(Wastegrd, 2002).
The climate varies somewhat between the study
areas, according to the 1961-1990 climate normals
(Alexandersson et al., 1991). Vrmland is characterised by relatively maritime climate in the south and
slightly more continental conditions in the north
(Vedin, 2004). In the investigated area, the mean
annual temperature is c. +4.6C and total precipitation is c. 700-800 mm/yr (Figure 7 a, b). In the
area of ngstugsmossen the mean annual temperature is c. +6.0 C, and a precipitation of c. 520 mm/
yr (Figure 7 c). In the area where ltabergsmossen
and Gullbergbymossen are located, the mean annual
temperature is c. +5.1 C and total precipitation is c.
555 mm/yr (Figure 7 d).

Methods
Field work
A large number of sites were selected for initial investigation. Preferably bogs would be geographically
well dened and contain distinct changes in peat humication. The stratigraphy should cover as much as
possible of the Holocene. The majority of the possible sites were discarded directly because they lacked
sufcient peat records, were severely drained/cut or
displayed indications of problematic stratigraphies
(e.g., hiatuses). For those bogs that were further investigated general stratigraphies were described. A
number of borings along transects were made using a
Russian peat corer to determine the stratigraphy for
each bog. In some cases a more detailed stratigraphy
was established (Paper IV). Sampling locations were
selected with respect to the established stratigraphies

12

15

40

-10

Temperature (C)

Precipitation (mm)

80

T: +4.8C
P: 794 mm

20

T: +5.1C
P: 555 mm

60

b) Djurskog
100

d) Svanberga

Temperature (C)

20

Figure 7. Precipitation (bars) and temperature (line) averages 1961-1990 (Alexandersson et al., 1991) for Torsby,
Djurskog (Vrmland), Svanberga (Uppland) and Herrberga (stergtland) the
meteorological stations situated closest
to the sites in this study. Annual mean
temperature (T) and total annual precipitation (P) is also shown.

Temperature (C)

T: +4.6C
P: 700 mm

Precipitation (mm)

80

a) Torsby

Temperature (C)

Precipitation (mm)

100

-10

and were in some cases also based on earlier investigations (J. Lundqvist, 1957, 1958a; Boygle, 1998). A
Russian peat corer of 1 m length and 5 cm diameter
was used for coring. In some cases the uppermost
part of the bogs was sampled with a half-cylindrical
PVC tube (50x10 cm; monoliths). All cores were
wrapped in plastic and transported back to Stockholm University for further analyses.
Laboratory analyses
Sample preparation
Peat cores and monoliths were stored in a cold room
(c. 5C) until subsampling was conducted. During
subsampling contiguous samples were cut out from
the peat cores and dried at 105C overnight. All
samples were ground and stored in 5 ml plastic containers.
Humication
The method to chemically analyse and determine the
decomposition of peat was originally developed by
Overbeck (according to Bahnson, 1968) and later
modied by Bahnson (1968). The method is based
on colorimetric determination of an alkaline extract
of the peat, where the alkaline absorbance is proportional to the amount of dissolved humic substances
(Aaby, 1976, 1986). Highly decomposed peat is usually dark brown/blackish whereas less decomposed
peat is light brown/yellowish in colour corresponding to the amount of extractable humic substances.
The methodology has been modied and improved
by Blackford and Chambers (1993), and this method
is now the standard used for determining peat decomposition. In this study a slightly modied version
has been used in order to better suite a large number
of samples. For the present study 0.1000.01 g dry
peat was dissolved in 25 ml 8% NaOH in 50 ml
plastic tubes, and boiled in a water bath at 95C for
1.5 hr.

The colour of climate: changes in peat decomposition as a proxy for climate change

was performed using a transfer function based on


modern testate amoebae assemblages (Woodland et
al., 1998; Schoning et al., 2005; Paper III).

Tephra

Figure 8. Distribution of 53 measurements of absorbance


on a standard sample consisting of Sphagnum peat, the
average value is 1.30 and standard deviation 0.04.

For the sequence from Stmyren (Paper I) the humic extracts were vacuum-ltered and diluted at the
dilution rate determined by Blackford and Chambers
(1993). In subsequent analyses, the boiled samples
were centrifuged and ltered and, in order to maintain the same dilution rate, 12.5 ml of the resulting
extract was diluted with distilled water to 100 ml.
Each sample was then measured three times in a UNICAM spectrophotometer at 540 nm and the average
was calculated and corrected to the normal-weight
of 0.1 g (Paper II). The reliability of the method was
tested by analysing a number of samples consisting
of one large batch of dried and ground Sphagnum
peat. The frequency of the measurements (Figure 8)
gives a distribution of 1.30.1 for the standard sample and approximates a normal distribution (Davis,
1986).
C/N
For carbon and nitrogen analyses 0.5-1 mg of dried
and ground peat was enclosed in metal capsules. The
analyses were performed with a Carlo Erba NC2500
analyzer connected, via a split interface to reduce
the gas volume, to a Finnigan MAT Delta plus mass
spectrometer at the Department of Geology and
Geochemistry, Stockholm University. The results are
shown as weight percentage and C/N ratio (Paper
III).
Testate amoebae
Analyses of testate amoebae assemblages were performed by Kristian Schoning (Gotland University)
following the sample preparation and taxonomy of
Charman et al. (2000). Contiguous 0.5-1 cm sub samples were obtained and thereafter mounted on slides.
A sum of 150-200 testate amoebae were counted in
each sample under a light microscope at a magnication of 400X. The reconstruction of the water tables

All cores except the lower parts of the sequences


from Kortlandamossen (c. 5000 to 10,000 cal yr BP)
were searched systematically for microscopic tephra
shards by Stefan Wastegrd (Stockholm University)
using the methods outlined by Pilcher et al. (1995).
Contiguous peat samples of 5-10 cm thickness were
combusted at 550C for 4 hr, washed in 10% HCl
and mounted for microscopy in Canada Balsam. For
those samples in which tephra shards were detected,
the equivalent sediment interval was re-sampled
using 1-2 cm slices, in order to determine more precisely the concentration and distribution of tephra
particles in each sample. The tephra shards were
prepared for geochemical analysis using Wavelength
Dispersive Spectrometry (WDS) at the Department
of Geology and Geophysics, Edinburgh University.
Samples were acid digested, following the procedure
outlined by Dugmore (1989). Results of the geochemical analyses are also presented in Wastegrd
(2005).
Radiocarbon dating
All accelerator mass spectrometer (AMS) radiocarbon dates were performed at ngstrm Laboratory, Uppsala University. Dating of the samples from
the Stmyren cores were performed on bulk peat,
whereas remaining dates were performed on plant
material >125 m and usually only leaves and stems
of Sphagnum were selected. If insufcient Sphagnum
was present or if the peat was to highly decomposed,
bulk material >125 m was used. Samples were dried
at 50-75C temperatures overnight and at 105C for
c. 3 hr.
Radiocarbon dates were calibrated with Calib 4.0
(Stuiver et al., 1998a) and OxCal 3.5 (Bronk Ramsey, 1995) using the INTCAL98 calibration curve.
All calibrated ages are shown with 2 condence intervals using the mid-point age (Table 3).
Chronology
The chronology for each peat bog was established
with the aid of the calibrated radiocarbon ages and
tephra horizons. For the sub-recent peat from ltabergsmossen and Gullbergbymossen (Paper III) spherical carbonaceous particles were also used (Schoning et al., 2005). The ages and geochemistry of the

13

Anders Borgmark

Table 3. AMS radiocarbon dates used in this study.


Site
Stmyren
Stmyren
Stmyren
Stmyren
Stmyren
Stmyren
Stmyren
Stmyren
Kortlandamossen
Kortlandamossen
Kortlandamossen
Kortlandamossen
Kortlandamossen
Kortlandamossen
Kortlandamossen
Kortlandamossen
Kortlandamossen
Kortlandamossen
Kortlandamossen
Fgelmossen
Fgelmossen
Fgelmossen
Fgelmossen
Fgelmossen
Fgelmossen
Fgelmossen
Fgelmossen
Fgelmossen
Fgelmossen
Fgelmossen
ngstugsmossen
ngstugsmossen

Core/depth
(cm)
Stm-20
Stm-38
Stm-54
Stm-60
Stm-76
Stm-99.5
Stm-150
Stm-187
Kort1-60
Kort1-140
Kort1-440
Kort1-540
Kort1-640
Kort2-110
Kort2-175
Kort2-320
Kort2-495
Kort2-550
Kort2-619
Fgm1-104
Fgm1-170
Fgm1-210
Fgm1-245
Fgm1-280
Fgm1-324
Fgm2-101
Fgm2-198
Fgm2-290
Fgm2-410
Fgm2-518
Ang-99
Ang-253

Laboratory
code
Ua-16209
Ua-16210
Ua-16211
Ua-16212
Ua-16213
Ua-16214
Ua-16215
Ua-16216
Ua-19026
Ua-19027
Ua-19028
Ua-19029
Ua-19030
Ua-22494
Ua-22495
Ua-23251
Ua-22496
Ua-23252
Ua-22497
Ua-22488
Ua-22489
Ua-23247
Ua-23679
Ua-23248
Ua-22490
Ua-22491
Ua-22492
Ua-23249
Ua-23250
Ua-22493
Ua-17865
Ua-17866

Peat component
dated
Bulk peat
Bulk peat
Wood
Bulk peat
Bulk peat
Bulk peat
Bulk peat
Bulk peat
Sphagnum
Sphagnum
Sphagnum
Sphagnum
Bulk>125m
Sphagnum
Sphagnum
Sphagnum
Sphagnum
Bulk>125m
Bulk>125m
Sphagnum
Sphagnum
Sphagnum
Sphagnum
Sphagnum
Sphagnum
Sphagnum
Sphagnum
Sphagnum
Sphagnum
Sphagnum
Sphagnum
Bulk>125m

mid Holocene tephra layers are well constrained


from several other investigations (e.g., Dugmore et
al., 1995; Pilcher et al., 1995; Larsen et al., 1999,
2001; Wastegrd et al., 2001; van den Bogaard et
al., 2002; van den Bogaard and Schmincke, 2002;
Bergman et al., 2004; Boygle, 2004) and provided
that the geochemical identication of the tephras is
robust, they can be used as xed age-depth points in
the age depth models. Age-depth relationships were
therefore dened using linear interpolation between
dated points, except at ltabergsmossen where a
second order regression was used by Schoning et al.
(2005).
The chronological uncertainty increases when
comparing sites and therefore full use of time parallel markers (i.e., tephras) should be made (Telford
et al., 2004). A linear interpolation age-depth curve
implies instantaneous shifts in the accumulation rate
at each radiocarbon date or tephra horizon, which
is extremely unlikely. However, whilst the curve is
forced to pass through the dates, it cannot deviate
too far from reality. Errors given by the uncertainties

14

14
C age
(yr BP)
75555
134055
2445100
248560
247555
348555
519075
719565
86550
142050
446565
610570
879085
85540
128040
190545
478545
698555
868570
85035
126535
207545
290040
329545
381040
60540
117540
192045
296045
424045
200565
7740100

2 calibrated
age (cal yr BP)
66697
125482
2483272
2551199
2549197
3743149
5961149
8017149
800120
133585
5090220
6980230
9875325
790120
1185105
1830120
5470140
7940130
9720190
790 120
1185105
2040120
3050160
3520120
4220140
60060
1100130
1850120
3140180
4740130
1971150
8647281

of radiocarbon dates i.e., calibrated 14C ages are unlikely to be the true date of that point, will be incorporated in the age-depth relationship when the curve
is forced to pass through every date (Telford et al.,
2004). The associated errors of calibrated ages are
not incorporated in linear interpolation age-depth
relationships either. The linear interpolation curve
has been used despite the incorporated errors in order to keep the age-depth relation of synchronous
tephra horizons xed.
Data analyses
In order to facilitate interpretation of the humication records, the data sets were subjected to standardisation and detrending (Davis, 1986; Chambers
et al., 1997; Chambers and Blackford, 2001). For
standardisation an average of 0 and a standard deviation of 1 were used in the equation:

Y=

The colour of climate: changes in peat decomposition as a proxy for climate change

Normalized abs

Raw humification

4
3

X
2
1

Step 1: Y=(X-)/

2
1

0
-1

Step 2: =0.0042Y-1.4044

Detrended abs

-2

Step 3: Y`=Y-

Y`

1
0
-1

50

100

150

200

250

300

350

400

450

500

550

600

650

Depth (cm)
Figure 9. Procedure for normalizing and detrending humication data. Step 1, normalization: From each absorbance
value (X) is the average () subtracted and the residual is divided with the standard deviation (). Step 2: The equation of
the linear regression is determined. Step 3, detrending: The results from the linear regression () are subtracted from the
normalized values (Y).

where X is the raw value, Y is the standardised value, is the average and is the standard deviation
of the data set (Figure 9). Detrending was performed
to minimize inuence from the decay that occurs in
the catotelm, which is extremely slow compared to
decomposition in the acrotelm (Clymo, 1984; Mauquoy and Barber, 1999; Belyea and Clymo, 2001),
and thus the subsequent trend is usually removed
(Mauquoy and Barber, 1999; Chiverrell, 2001). To
remove the trend a linear regression equation was
used (Figure 9), where the resulting value from the
equation at every data point () is subtracted from
each corresponding standardised value (Y) resulting
in a detrended value (Y`=Y-).
Spectral analysis has become an important tool for
interpretation of palaeo-climatic data, particularly
for identifying traces of external factors such as orbitally induced climatic oscillations (e.g., Milankovich cycles) or solar forcing. The so called quasiperiodic climate processes have also gained more
attention lately (Burroughs, 2001). These processes
are not strictly periodic and often are the result of
complicated and only partially understood interactions within Earths climate system e.g., the North

Atlantic Oscillation (NAO) and El Nio/ENSO


(Burroughs, 2001; Marshall et al., 2001; Lindeberg,
2002). Spectral analysis in Paper II was conducted
with the aid of a Lomb-Scargle Fourier transform,
which can handle unevenly spaced data without using interpolation (Schulz and Stattegger, 1997). The
analyses were performed with the program SPECTRUM (Schulz and Stattegger, 1997) and the estimations of the red-noise (i.e., noise that has more power
at lower frequencies; Lindeberg, 2002) spectra were
done with the program REDFIT (Schulz and Mudelsee, 2002), which constructs an estimation of the
rst-order autoregressive parameter directly from an
unevenly spaced time series. This makes it possible
to test the null hypothesis that the spectrum obtained
from the time series originates from a rst-order autoregressive process i.e., noise.
A composite record was established by compiling data from ve sequences covering the last 4000
years. This time frame was divided into even 50-yr
intervals where the closest humication values from
each separate sequence were positioned. At those
levels where three or more standardised humication values were present an average was calculated.

15

Anders Borgmark

If changes in humication are at random, then successive averages should not have the same value, nor
should they display a trend. When applying this relatively crude methodology a number of errors could
affect the outcome. The composite record does, however, provide a broad picture of regional changes in
bog hydrology.

Results
Summary of paper I
Gunnarson, B. E., Borgmark, A. and Wastegrd,
S. 2003: Holocene humidity uctuations in Sweden
inferred from dendrochronology and peat stratigraphy. Boreas 32, 347-360.
This paper tests the hypothesis that past hydrological changes are not only locally induced but can
be found also over larger regions such as central
Sweden. Tree ring and humication data covering
the last 7 000 years were analysed in order to identify and compare humidity changes in Fennoscandia.
Subfossil woods of Scots pine (Pinus sylvestris) found
in lakes around Lake Hckren, Jmtland (Figure 1),
were used to provide a dendrochronological data
set. The dendrochronology was standardised, using
a RCA method, and ltered to show low frequency
variations that could be compared to the peat humication record from Stmyren, Vrmland (Figure 1).
The chronology of the peat sequence was established
with aid of six AMS radiocarbon dates and three tephra horizons. One additional tephra that was previously not described was also found. For peat humication measurements, a modied standard method
was used.
Interpretation of the regeneration of pine, peat humication and tree ring growth tentatively indicate
synchronous periods of drier climate c. 6850-6750,
4350-4150, 4050-3750, 3450-3050, 1900-1750,
1550-1350 and 600-450 cal yr BP. Possible wetter
periods were inferred at 5550-5350, 5150-4850,
4150-4050, 3650-3450, 3050-2850, 2050-1900,
1750-1550, 1200-1050 and 400-250 cal yr BP. Note
that in the paper all ages are shown as BC/AD. The
wet and dry periods revealed by the tree ring and
peat stratigraphy data indicate considerable humidity changes during the Holocene.
Bjrn Gunnarson conducted the dendrochronological analyses and interpretations, Anders Borgmark
performed peat humication analyses and interpretations and Stefan Wastegrd analysed the cores for
tephra. All authors are equally responsible for the
discussion and conclusions.

16

Summary of paper II
Borgmark, A. 2005: Holocene climate variability and
periodicities in south-central Sweden, as interpreted
from peat humication analysis. The Holocene 15,
387-395.
This paper deals with analyses of peat sequences
from Stmyren and Kortlandamossen in Vrmland
(Figure 1). The focus of the study was to identify
synchronous shifts between low and high humied
peat in two ombrotrophic peatlands and to evaluate the potential of spectral analysis as a tool for
analysing climate data derived from peat humication variations. Both bogs were probably formed by
paludication in the early Holocene. Chronologies
were established with AMS radiocarbon dating and
tephra horizons. Stmyren covers the last 8000 years
and Kortlandamossen was initiated nearly 10,000
years ago according to the radiocarbon dates. The
humication data was interpreted in terms of wet
and dry periods and the data sets were also subjected to spectral analysis, using the Lomb-Scargle
Fourier transform. One advantage of this method
is that the transform function can analyse unevenly
spaced data, a common characteristic of palaeoclimatic time series. The resulting spectra have to be
interpreted with caution, but application of an AR1
model, which gives a null hypothesis of the possibility that the spectra only show red noise, strengthens
the interpretation that there are signicant periodicities in the data. Frequencies that have a wavelength
less than twice the spacing between sample points
cannot be detected, since the highest frequency that
can be estimated is the Nyquist frequency, which is
the wavelength exactly twice the distance between
successive observations. Each sample represents an
age of c. 50 years in Kortlandamossen and c. 75
years in Stmyren, and thus frequencies shorter than
100-150 years are impossible to identify. In the low
frequency area, the inuence of red noise increases,
indicated by a rise in the AR1 model. Therefore,
frequencies longer than c. 500 years are difcult to
identify, leaving a window of c. 100-500 years where
periodicities can be recognized.
The results indicate that there are a number of
more or less synchronous shifts in hydrology in the
two bogs. The most pronounced shifts occur at c.
7500 cal yr BP (wet-shift), 4500 cal yr BP (dry-shift),
4300-4200 cal yr BP (wet-shift), 3700-3500 cal yr
BP (wet-shift), 3250 cal yr BP (dry-shift) and 2250
cal yr BP (wet-shift).
A periodicity of around 250 yr is evident in the
spectral analysis of both cores, and periodicities of
the same magnitude have been reported from other
proxies as well as from other peat based studies. A

The colour of climate: changes in peat decomposition as a proxy for climate change

periodicity of c. 200 yr, possibly corresponding to


the Suess solar cycle, is found in the spectral data
from Stmyren and a periodicity of around 400 yr
that could be of solar origin is found at Kortlandamossen. The results of spectral analysis have to be
interpreted with caution because of the level of noise,
particularly frequencies longer than c. 500 yr should
be interpreted carefully when data of this type and
resolution are at hand.
Anders Borgmark performed all analyses and interpretations and is responsible for discussions and
conclusions. The tephrochronological work was performed by Stefan Wastegrd and geochemical data is
presented in Wastegrd (2005).
Summary of paper III
Borgmark, A. and Schoning, K. in press: A comparative study of peat proxies from two eastern central
Swedish bogs and their relation to meteorological
data. Journal of Quaternary Science.
Peat humication, C/N and testate amoebae assemblages were analysed on a high resolution peat
sequence from ltabergsmossen spanning the last
150 yr, and on a shorter sequence from Gullbergbymossen spanning the last 60 yr, testate amoebae and
peat humication were analysed (Figure karta).
The physical and chemical proxies were correlated
to each other and compared to the testate amoebae
assemblage inferred water table and to meteorological data from Uppsala. On the sequence from
Gullbergbymossen only the uppermost part could
be used (corresponding to ca 1940-2000) because
of a possible hiatus in the peat sequence. The chronologies were established by a combination of SCPs,
radiocarbon dates and the Askja AD 1875 tephra
(Schoning et al., 2005). The age-depth relationships
imply a peat accumulation at a rate of 2.5-5 mm/yr
at both sites, providing an excellent opportunity to
compare the peat proxies with available temperature
and precipitation data for the last 150 years.
Peat humication was measured using the modied standard method, and carbon and nitrogen content were analysed on the same samples. The testate
amoebae were counted according to standard methodology.
Peat humication and the inferred water table
show signicant changes with the same trends in
both records. This indicates that changes in water
table are driven by external forcing. It was not possible to measure to a statistically signicant level the
responses or the lag of the physical/chemical parameters (peat humication and concentration of carbon

and nitrogen) versus the meteorological/biological


(testate amoebae assemblages) parameters in this
study.
The results indicate, however, a lag between the
biological parameter and the physical/chemical parameters. This lag is likely due to the decay of deeper
layers in the peat during drier periods. The term secondary decomposition is used for the process of further decay of previously deposited peat due to a lowering of the water table. This process has implications
for the interpretation of multi- and single-proxy data
from ombrotrophic mires, especially when interpreting recent and/or high-resolution data. Biological
proxies e.g., plant macrofossils and testate amoebae,
reect the moisture conditions at the time when the
plants and amoebae lived. Peat humication and C/N
ratio on the other hand reect conditions during the
decay-process of the organic material i.e., processes
that inuence previously deposited material.
We conclude that it is necessary to perform highresolution multiproxy studies of peat sequences with
good temporal resolution to be able to interpret relationships and leads/lags between biological and
physical/chemical proxies.
Anders Borgmark prepared the samples for C/N
analysis and did the peat humication analyses and
the interpretations of these two proxies. Kristian
Schoning analysed and interpreted the testate amoebae assemblages and provided the chronologies.
Both authors are equally responsible for the discussion and conclusions.
Summary of paper IV
Borgmark, A. and Wastegrd S.: Regional and local patterns of peat humication in three raised peat
bogs in Vrmland, south-central Sweden. Manuscript.
Peat humication data from three bogs in Vrmland, south-central Sweden, is presented, with replicate sequences from two of the bogs. The degree
of peat decomposition was used to infer past bog
surface wetness. A comparison between the results
from the peat humication analyses and a number
of different records of Holocene climate change was
conducted, in order to interpret similarities and dissimilarities and their possible causes.
Tephra horizons and a total of 30 AMS radiocarbon
dates constrained the chronologies (Table 3). Tephra
from ve different eruptions were found in the cores
and geochemically ngerprinted. The Kebister and
Hekla-3 and 4 tephras were found in several cores,
whereas Askja 1875 and the Stmyren tephra were

17

Anders Borgmark

Stefan Wastegrd searched the peat sequences for


tephra horizons, conducted the tephra analysis and
interpreted the geochemical data. Anders Borgmark
performed peat humication analyses and interpretations of the humication records. The interpretation of climatic changes has been performed by both
authors.

18

Additional results
Cores were collected from three additional bogs in
Vrmland (Figure 1) and partly analysed for humication and tephrochronology. However, the data has
not been used because of indications of anthropogenic disturbances at the sites. Brrudsmossen had
earlier been cut and an initial analysis of the peat humication revealed a disturbed record. Mosstakanmossen was cored, but when visited for resampling it
was drained and peat-cutting had commenced. The
bog was therefore no longer suitable for the replicate
core research it was intended for. Klaxsjmossen had
also been extensively cut and, based on the disturbed
peat humication record shown at Brrudsmossen,
the bog was excluded.
The peat humication record (Figure 10) from
ngstugsmossen, stergtland (Figure 1), supports
the conclusion of Granlund (1932) that very few
shifts in peat humication i.e., recurrence surfaces,
are found in this part of Sweden. Two radiocarbon
dates from the prole (Table 3) indicate a total age of
c. 8700 cal yr for the peat. Transitions to less humied peat can be found at c. 3000 and c. 600 cal yr
BP, with the former transition being gradual and the
latter relatively abrupt.

10
20
30
40
50
60
70
80
90

100
Depth (cm)

only found in one core, respectively. The tephras are


time-synchronous markers, providing excellent opportunities for core correlation, and to evaluate if
changes in surface wetness were synchronous within
a single peatland.
The replicate cores from Fgelmossen display similar patterns of changes in surface wetness at times,
which is also the case at Kortlandamossen where the
sequences, covering almost 10,000 years, seem to be
somewhat offset to one another. The mid Holocene
cooling (6000-5000 cal yr BP) is not directly evident in the data from Stmyren. In Kortlandamossen a change to more unstable conditions starts at ca
5500-5000 cal yr BP, with a wet-shift present in both
sequences. Both sequences from Kortlandamossen
display wet-shifts within this period, but they do also
have indications of drier conditions and dry-shifts.
The humication record of Stmyren indicates high
to medium water table. This period has, however,
been reported as a dry period in Fennoscandia which
in combination with a general cooling possibly could
result in this type of humication record.
All three bogs indicate a change to wetter conditions between 4500 and 4000 cal yr BP, as many other records do across the North Atlantic region. Between 3700 and 3200 cal yr BP, wet-shifts occurred
in the three Vrmland bogs as well as in many other
European bogs, simultaneous with decreasing temperatures, registered in e.g., the Greenland ice cores
(GISP2) and Norwegian speleothems. The 2800 cal
yr BP climatic change is evident in all bogs. At Kortlandamossen, this cooling may have caused a hydrological threshold to be crossed, leading to generally
lower decomposition of the peat i.e., prevailing wet
surface conditions. A large number of wet-shifts are
present during the last 2500 yrs at European bogs;
around 1600 cal yr BP, several records indicate a climate change that is visible as wet-shifts in three of
the analysed sequences.
The composite record, covering the last 4000 yr,
shows distinct wet-shifts at c. 3800-3300, 31002800 and 1700-900 cal yr BP, which correlate well
with many other climate records. The composite
record could be a powerful tool in interpreting regional changes in peat hydrology caused by climatic
variation.

110
120
130
140

150
160
170
180
190
200
210
220
230
240
250

260
1 2 3 4 5 6
Absorbance

Figure 10. Raw humication (absorbance) values from


ngstugsmossen, stergtland. The positions of the two
AMS-radiocarbon dates (Table 3) and the Hekla-4 tephra
horizon are also shown.

The colour of climate: changes in peat decomposition as a proxy for climate change

Discussion
Local correlation of humication records
Due to the nature of raised bogs one could not expect
to obtain stratigraphical concordance over the entire
bog, since natural processes at different scales inuence peat formation, accumulation and decomposition (Figure 5). There are several reports of differing
results from replicate records (e.g., Caseldine et al.,
1998; Mauquoy et al., 2002), which emphasize the
need for replicate records. However, it has also been
noted that replicability improves over time, and that
this improvement could be explained by bog expansion (laterally and vertically) leading to surface homogeneity, which in turn decreases the inuence of
local hydrological changes on the bog surface (Charman et al., 1999).
At the sites Kortlandamossen and Fgelmossen
replicate peat sequences were analysed, providing
an opportunity to evaluate at which scales changes
occur and which processes affect surface wetness at
meso and macroscale.
Since the replicate records of bog hydrological
changes were obtained at distances shorter than
1000 m but more that 100 m, mesoscale (Figure 5)
autogenic processes cannot be ruled out. However,
mesoscale changes in surface wetness are quite likely
to be linked to climate (Charman, 2002), and therefore changes that occur more or less simultaneously
are likely to be of climatic origin.
At both Kortlandamossen and Fgelmossen a
number of changes fall within this criterion, but usually the magnitude and/or the timing differ to some
degree. The differences could presumably be attributed to microscale and some mesoscale processes,
such as microtopography, microclimate, succession
and vegetation and the hydrological thresholds that
result from them (e.g., Barber, 1981; Charman,
2002). However, the presence of similar trends in the
replicate cores implies that external forcing caused


cal yr BP

Figure 11. The ve recurrence surfaces from Granlund


(1932) plotted together with the peat composite record
from Vrmland (average of humication index; Paper IV).
Age intervals of recurrence surfaces are Granlunds age
(transformed to cal yr BP) 100 yr.

much of the changes in bog surface wetness. The


four records also imply that replicability increases
over time in both Fgelmossen and Kortlandamossen (cf., Charman et al., 1999).
Regional correlation of humication records
It is evident in this study that periods of mainly wetshifts are temporally clustered in bogs on a regional
scale. This has also been shown in several other studies (e.g., Anderson, 1998; Blackford, 1998; Barber et
al., 2000, 2004; Hughes et al., 2000; Mauquoy and
Barber, 2002; Langdon et al., 2003).
In paper IV, the regional pattern of changes in peat
humication and the composite record (Paper IV:
Figure 6) indicates periods of wetter surface conditions between c. 3700-3200 and 1700-900 cal yr
BP and periods of relatively dry conditions between
2600-1900 and 800-300 cal yr BP. Additionally, a
very distinct wet-shift is evident at c. 3000-2800 cal
yr BP. Between c. 500 cal yr BP and the present, the
composite record seems to indicate increasing surface wetness on the investigated bogs. These periods
of relatively good concordance give an indication of
the resolution at which regional patterns can be recognized, given the constraining factors of the composite record (Paper IV; nearest 50-yr sample, time
resolution and age-depth relationships). By improving the techniques used for compiling records, it is
likely that more information could be obtained and
this type of study might then be a complement to
multi-proxy studies of one or two sequences.
Despite the uncertain ages of Granlunds (1932)
recurrence surfaces (Table 1) there seems to be a fair
agreement between the main shifts in peat humication in Vrmland and Granlunds recurrence surfaces
(Paper I; Paper II). A diagram with Granlunds recurrence surfaces I-V superimposed on the regional humication record from Vrmland (Figure 11) shows
that there are indications of wet conditions close to
the four recurrence surfaces covered by the time period of the composite record. This implies that the
scheme set up by Granlund is principally correct,
however, since Granlunds ages of these shifts are
very uncertain should the term recurrence surface
perhaps be used only in its conceptual sense and focus should not be on Granlunds interpreted ages of
the recurrence surfaces.
Comparison to other climate data
As outlined in Paper I several synchronous periods
of drier/wetter conditions can be identied when
comparing peat humication data and lake-level

19

Anders Borgmark

uctuations inferred from the temporal distribution of Scots pine, even though the records originate
from different regions in Sweden. Synchronous patterns of hydrological change between different types
of archives located close to each other have been
found by e.g., Baker et al. (1999) and Charman et
al. (2001). Many climate archives in the North Atlantic region display a pattern of coherent climatic
events (Figure 6), and on a broad scale many shifts in
bog surface wetness appear at these times of climatic
change. Three periods of inferred changes in bog surface wetness are detectable in the composite record
at c. 3700-3200, 3000-2800 and 1700-900 cal yr
BP (Figure 6 h) and their correspondence to other
palaeoclimate records on different spatial scales are
discussed further here. The period of wetter conditions between 3700 and 3200 cal yr BP can be found
in all analysed sequences presented in Paper IV. A
transition to wetter conditions during this period has
been found in lake records from southern Sweden
(e.g., Digerfeldt, 1988; Hammarlund et al., 2003;
Jessen et al., 2005), and a cooling and an increase
of winter storminess during this period are inferred
from many other Scandinavian records (Lauritzen
and Lundberg, 1999a; Snowball et al., 1999; Bjrck
and Clemmensen, 2004). In north-western Europe
many wet-shifts are registered in bogs (Figure 6 f) at
this time (Nilssen and Vorren, 1991; Anderson et al.,
1998; Hughes et al., 2000; Barber et al., 2004) and
a period of increased lake-levels has also been reported in central Europe (Magny, 2004). On a larger
scale has an increase in Na+ in the GISP2 record been
detected (Figure 6 b), indicating a deeper Icelandic
Low (Mayewski et al., 1997), however, there are
comparably few other records from the North Atlantic region that indicate a major climatic change
at this time. The implication of this could be that
this climate change consisted of a change primary in
humidity and that any change in temperature was of
minor importance.
A wet-shift in the composite record at around
3000 cal yr BP (Figure 6 h) could be traced in all the
investigated bogs from Vrmland (Paper IV; Figure
6). This wet-shift begins slightly before a drastic increase in 14C at c. 2800 cal yr BP (Figure 6 g) that
has been linked to a global climate change (e.g., van
Geel et al., 1998, 2000; Viau et al., 2002; Blaauw
et al., 2004). There are several indications of a climate change in Scandinavia at this time (e.g., Denton and Karln, 1973; Karln and Kuylenstierna,
1996; Lauritzen and Lundberg, 1999a; Rosqvist et
al., 2004) and a number of palaeorecords indicate a
lowering in temperatures also on a larger scale at this
time (e.g., OBrien et al., 1995; Bond et al., 1997,

20

2001; Mayewski et al., 1997, 2004). Throughout


north-western Europe many wet-shifts (Figure 6 f)
are reported in bogs at this time (Nilssen and Vorren,
1991; Kilian et al., 1995; Hughes et al., 2000;
Langdon et al., 2003; Barber et al., 2004; Blaauw et
al., 2004) and together with increased lake-levels in
the Jura mountains (Magny, 2004) this implies that
this was a period of increased humidity in Europe.
The lowering of temperatures together with the
decrease in solar activity could have led to a more
humid climate, perhaps due to decreased evaporation which increased surface wetness conditions on
north-western European bogs.
The period between c. 1700 and 900 cal yr BP is
characterised by mainly wet conditions in Vrmland according to the composite record. Many other
European bogs seem to have reacted to an increase
in surface wetness during this time as well (Figure
6 f)(Nilssen and Vorren, 1991; Hughes et al., 2000;
Chiverrell, 2001; Langdon et al., 2003; Barber et
al., 2004). An increase in lake-levels in southern
Sweden (Digerfeldt, 1988), decreasing temperatures
in Norway (Lauritzen and Lundberg, 1999a) and
glacier advances in Scandinavia (Denton and Karln,
1973; Karln and Kuylenstierna, 1996; Matthews et
al., 2000; Rosqvist et al., 2004) indicates that there
was a regional shift in climate at this time. Several
palaeorecords indicate a climatic transition around
1500 cal yr BP in Europe and the North Atlantic
region (Figure 6 a, b, d, e)(Mayewski et al., 1997,
2004; Bond et al., 2001; Magny, 2004).
The three periods of increased bog surface wetness
deduced from the composite record discussed here
are all in concordance with climate changes found in
other records and in other regions, especially within
the North Atlantic region. The climate shift c. 30002800 cal yr BP is registered in several records around
north-west Europe, this climate change has also been
found in palaeorecords elsewhere in the world and
a decrease in solar irradiation seem to have affected
the climate at this time. Many palaeoclimate records
also display a climate change starting c. 1700 cal yr
BP. In this case, however, there is no direct evidence
of a change in solar activity and indications of a
global change are few (e.g., Mayewski et al., 2004).
Finally, few archives other than bogs and lakes indicate a climate change around 3700 cal yr BP around
the North Atlantic region and few records indicates
cooler conditions during this time. Despite the Norwegian speleothem record and indications of glacier
advances in Scandinavia (Lauritzen and Lundberg,
1999a; Snowball et al., 1999) this climate change
seems to have been more a change of humidity than
in temperature in Scandinavia. From this it can be

The colour of climate: changes in peat decomposition as a proxy for climate change

argued that bogs react to climate changes at different spatial scales and to different types of changes
(i.e., temperature and humidity). It is clear that the
bogs investigated in this study have reacted to the
larger Holocene climatic shifts throughout the North
Atlantic region. However, the peat decomposition
processes react also to changes in local conditions
that to some extent can bias the interpretation. In
this respect a composite record could be useful in
order to lter some of the locally induced changes in
bog surface wetness.
Causes of hydrological shifts and periodicities in
bogs
Since both external and internal forcing cause changes in climate (Emeis and Dawson, 2003), variation
in solar activity is a potential forcing factor for climatic variability on decadal to millennial timescales
(Chambers et al., 1999; Bond et al., 2001). Increases
in 14C (Figure 6 g) at c. 5500 and c. 2800 cal yr BP
indicate a decrease in solar irradiation at these times,
which could have caused climate changes (van Geel
et al., 1998, 2000; Mayewski et al., 2004). The decrease in solar activity at 5500 cal yr BP could have
caused the wet-shift present in Kortlandamossen
(Paper IV; Figure 6) and the distinct shift to low solar
activity at 2800 cal yr BP is recorded in all investigated sites in Vrmland as a period of wet conditions
(Paper IV; Figure 6).
Three different cycles were found in the humication data from Stmyren and Kortlandamossen: c.
250 yr (both bogs), c. 200 yr (Stmyren) and c. 400
yr (Kortlandamossen). Periodicities of 25025 yr
were identied in distal glaciolacustrine sediments in
southern Norway (Matthews et al., 2000), in sediments in the Gotland Basin (Kunzendorf and Larsen,
2002) and at Draved Mose in Denmark (Aaby,
1976). Both the c. 200 yr and c. 400 yr cycle have
been found in other studies, and could be of solar
origin (Wijmstra et al., 1984; Stuiver and Braziunas,
1989; Chambers et al., 1997; Dean and Schwalb,
2000; Chambers and Blackford, 2001; Dean et al.,
2002; Clilverd et al., 2003; Yu, 2003).
The changes and periodicities discussed might
result from changes in solar activity in a complex
cryosphere-ocean-atmosphere system and implies
that solar activity plays an important role in climate
oscillations over the North Atlantic region (e.g., Karln and Kuylenstierna, 1996; Chambers et al., 1999;
van Geel et al., 1999; Bond et al., 2001; Chambers
and Blackford, 2001; Carslaw et al., 2002; Frhlich
and Lean, 2002; Clilverd et al., 2003; Blaauw et al.,
2004; Magny, 2004).

In the two records (ltabergsmossen and Gullbergbymossen; Paper III) that have been compared
to meteorological records decreased precipitation
could be determined as a cause for lowering of the
water tables, however, higher summer temperatures
at the initiation of the dry period could also have
inuenced the bog surface wetness.
The results from the ve analysed peat sequences
from Vrmland show that shifts in bog surface wetness are present at times of decreased solar activity as
well as when there is no indication of such a change.
Additionally do the results from all analysed bogs
in this study show that changes to wetter conditions
on bogs occur when either temperature decreases
or precipitation increases, or if they both change at
the same time. This implies that no explicit forcing
mechanism, other than climate, can be determined
in this study.
Problems and potentials with peat humication
The normal distribution of the frequency of the
standard sample (Figure 8) implies that the method of chemical extraction of humic substances and
the analytical protocol used are acceptable. However, this does not take into account the problems
of NaOH alteration of the humic acids from high
to low molecular weight (Caseldine et al., 2000).
The problems of poorly understood processes in the
standard methodology need attention, but nevertheless a thoroughly applied procedure can provide a
palaeohydrological proxy (Caseldine et al., 2000).
Transformation of the measured absorbance to percentage humication should be avoided, because
of the variety of humic acids produced by different peats under different conditions (Blackford and
Chambers, 1993). In the present study most humication records have been normalised, both to facilitate comparison between different sequences with
sometimes large differences in mean humication
levels, and to reduce the risk of comparing levels of
humication measurements as degree of decomposition which may lead to misinterpretations.
The process of secondary decomposition (Tipping,
1995) described in Paper III has great implications
for palaeoenvironmental interpretations of humication records. The onset of the effects of drier conditions could appear to start earlier (deeper) than
the actual hydrological shift. However, this process
can be discriminated by the compaction that usually
takes place in longer sequences. When analysing long
series, even in contiguously sampled peat sequences,
the temporal resolution could be too low to detect
the process of secondary decomposition. However,

21

Anders Borgmark

when performing high-resolution analyses of peatsequences (i.e., sub-recent peat), the process of secondary decomposition should be taken into account
when interpreting the data.
Perhaps the most complicating factor with peat humication as a proxy for past climate change is that
the method only produces a relative measurement of
change, it is not possible to estimate the magnitude
of change or to say that a certain degree of humication equals a certain width of the acrotelm. One way
to overcome this problem might be to use e.g., testate amoebae assemblages for which transfer functions have been developed, which makes it possible
to obtain quantitative estimates of peat moisture
and water table depth (Warner and Charman, 1994;
Charman and Warner, 1997; Woodland et al., 1998;
Charman et al., 2000; Charman, 2001; Mitchell et
al., 2001; Booth, 2002). When comparing peat humication records and water tables inferred from testate amoebae assemblages from a sequence, it might
be possible to transform the humication data to
bog surface wetness using a two-step transfer function. A function of this kind could be set up by using
a relatively small number of samples from which testate amoebae assemblages are determined. The testate amoebae transfer function is then used to infer
a water table for these assemblages, the results are
transferred to the matching peat humication values
and from that surface wetness could be inferred on
the entire sequence.
Due to the nature of the decomposition process,
this two-step transfer function is valid only on the
actual sequence analysed. However, usage of the relatively fast and unproblematic determination of peat
humication, together with a small number of analyses of testate amoebae assemblages, would make it
possible to retrieve a better result than if only one
method is used and less time consuming than if both
methods were used separately (i.e., more samples of
testate amoebae analysed).

Conclusions
Comparison of replicate sequences from one bog
gives an idea of the type and scale of active hydrological processes i.e., if they are autogenic (microand mesoscale) or allogenic (meso- and macroscale).
Comparison and correlation of sequences from several bogs within a specic area provides indication
of the macroscale processes, of which climate is the
most important factor.
Changes in peat bog surface wetness can be correlated with other Fennoscandian palaeoclimate
records as well as climate shifts registered across the

22

North Atlantic region. This implies that the peat humication displays changes in bog hydrological conditions due to variations in the climate in a scale of at
least north-western Europe and probably the North
Atlantic region. Increased surface wetness has been
inferred from a composite record, c. 3700-3200,
3000-2800 and 1700-900 cal yr BP.
Cyclic patterns of inferred bog surface wetness
have been found with frequencies of 200, 250 and
400 yr. These periodicities and the wet-shifts at c.
5500 and 2800 cal yr BP correlate to variations in
sun irradiance and it is plausible that this external
forcing caused the shifts in climate. In this study it
was evident that, given the physical properties of the
data, frequencies longer than c. 500 yr could be the
result of a rst-order autoregressive process implying
that great caution should be taken when interpreting
this type of data.
Peat humication reects shifts in the surface
wetness of raised peat bogs. The processes of decomposition are, however, not fully understood. Secondary decomposition could result in a lag between
peat humication and biological proxies, especially
when analysing high-resolution records. Due to the
complexity of peat decomposition processes it is important to interpret humication data carefully, with
special emphasis on the amplitude needed in the humication record for interpretation of possible climate driven changes.
In order to better assess the relationship between
precipitation and temperature and their resulting inuence on bog surface wetness more studies of the
processes involved in peat decomposition are necessary. Usage of multiproxy high-resolution studies of
sub-recent peat that could be correlated to known
climatic conditions is probably the best way to study
decomposition processes.
Chemical extraction of humic substances, despite
limitations in the extraction process, is a straightforward method that produces replicable results. To
delimit the number of errors, well-dened mediumsized ombrotrophic bogs without signs of drainage
and/or peat cutting should be used in this type of
study.

The colour of climate: changes in peat decomposition as a proxy for climate change

Acknowledgements
Under dessa r har min familj varit ett stort std,
ven om de inte alltid har frsttt vad jag sysslat med
har Mamma, Pappa, Ullis, Kattis och Fredde alltid
funnits dr.
Utan Elin och Axel hade livet varit halvt och bra
mycket trkigare. Mitt allra strsta och djupaste
tack gr till Johanna, som inte bara har sttt ut med
mig under den hr tiden, utan ocks stttat mig nr
jag behvt det som mest.
Stefan Wastegrd initiated this project and together with Ulrik Kautsky (SKB) we formulated the main
focus of the work. Stefan has, as project leader and
main supervisor, been there with a good balance between whip and carrot. Ann-Marie Robertsson, my
other supervisor, has always kindly read my texts. I
could not have done this without the two of you.
Jan Lundqvist has always been supportive of my
work, and his stories about the good old days will
always be remembered - and I hope that some of the
mess in Vrmland has been cleaned up now!
Fieldwork and the majority of the radiocarbon
dates were funded by the Swedish Nuclear Fuel and
Waste Management Co (SKB) and by scholarships
from the foundations of Helge Ax:son Johnson and
Gerard De Geer. For the rst year of PhD studies I
received a scholarship from Gustaf and Ellen Kobbs
Foundation. The remaining years were funded by
Stockholm University and a grant from The Swedish Natural Science Research Council to Stefan
Wastegrd. Financial support has also been provided
by The Swedish Foundation for International Cooperation in Research and Higher Education (STINT),
Wallenbergs jubileumsfond, C F Liljevalch J:ors resestipendium, Carl Mannerfelts Foundation and The
Swedish Natural Science Research Council (grant to
Stefan Wastegrd).
This thesis has beneted greatly from the comments, questions and criticism from Ninis Rosqvist,
Jan Risberg and Barbara Wohlfarth and Im grateful
to you all. Terri Lacourse made a huge effort in correcting the language of the thesis.
My brother in arms Kristian Schoning has been a
great support while driving around in the forests in
search for the optimal bog, and coring the potential
ones. Since we have yet to nd it, there is still work
to do
Many people that have contributed with bits and
pieces over the years and you are all acknowledged.
However, I have to mention some of you; Greger
Lindeberg introduced me to the magical world of
frequency analysis and showed how dark that magic
can be. Anders Moberg has been very helpful with

answers to questions about time-series and statistics.


Ann Karlsson helped with futher development of the
humic acid extraction methodology. Ann, and later
Sven Karlsson, have been invaluable in the lab. Katarina Borgmark nished the last batch of humication
measurements and eased my burden.
As a source of knowledge about almost everything
from H. P. Lovecraft to Return to Castle Wolfenstein
has Jens Heimdahl been more than just a colleague
and we have had lots of fun together during the years.
I have had several fruitful and enjoyable discussions
about peat, climate and everything in-between with
Jonas Bergman.
Keith Barber kindly arranged my visit at the Department of Geography, University of Southampton
(funded by STINT), he and Paul Hughes, Pete Langdon, Antony Blundell, Paul Coombes and the rest
of the gang made my time there (and at the pubs!) a
memorable and learning experience.
Per, Greger, Anna, Kristian, Gun, Tiit, Jens, Martina and all the other past and present PhD students
in Quaternary Geology have made me proud of being part of this group, as everyone at the Department
of Physical Geography and Quaternary Geology has.
The innebandy-players took my mind of science
for at least one hour every weekit has been great
fun!

References
Aaby, B. 1976. Cyclic climatic variations in climate over
the past 5,500 yr. reected in raised bogs. Nature 263,
281-284.
Aaby, B. 1986. Palaeoecological studies of mires. In Handbook of Holocene Palaeoecology and Palaeohydrology.
B. E. Berglund (Ed.). John Wiley & Sons Ltd. 145-164.
Aaby, B. and Tauber, H. 1974. Rates of peat formation in
relation to degree of humication and local environment,
as shown by studies of a raised bog in Denmark. Boreas
4, 1-17.
Aario, L. 1932. Panzentopographische und paleogeographische Mooruntersuchungen in N-Satakunta.
Fennia 55, 1-179.
Alexandersson, H. and Andersson, T. 1995. Nederbrd
och ska. In Klimat, Sjar och Vattendrag. B. Raab and
H. Vedin (Eds.), Sveriges Nationalatlas. Bokfrlaget Bra
Bcker, Hgans. 76-90.
Alexandersson, H., Karlstrm, C. and Larsson-McCann,
S. 1991. Temperature and precipitation in Sweden
1961-1990. Reference normals. SMHI Meteorologi 81;
88 pp.

23

Anders Borgmark

Alley, R. B., Mayewski, P. A., Sowers, T., Stuiver, M.,


Taylor, K. C. and Clark, P. U. 1997. Holocene climatic
instability: A prominent, widespread event 8200 yr ago.
Geology 25, 483-486.

1999. Proxy records of climate change in the UK over


the last two millennia: documented change and sedimentary records from lakes and bogs. Journal of the Geological Society, London 156, 369-380.

Alverson, K., Bradley, R., Briffa, K., Cole, J., Hughes, M.,
Larocque, I., Pedersen, T., Thompson, L. and Tudhope,
S. 2001. A Global Paleoclimate Observing System.
Science 293, 47-48.

Barber, K. E., Chambers, F. M. and Maddy, D. 2004. Late


Holocene climatic history of northern Germany and
Denmark: peat macrofossil investigations at Dosenmoor,
Schleswig-Holstein, and Svanemose, Jutland. Boreas 33,
132-144.

Anderson, D. E. 1998. A reconstruction of Holocene


climatic changes from peat bogs in north-west Scotland.
Boreas 27, 208-224.
Anderson, D. E., Binney, H. A. and Smith, M. A. 1998.
Evidence for abrupt climatic change in northern
Scotland between 3900 and 3500 calendar years BP.
The Holocene 8, 97-103.
Backus, I. 1990. The Cyclic Regeneration on Bogs - a
Hypothesis that became an Established Truth. Striae 31,
33-35.

Barber, K. E., Maddy, D., Rose, N., Stevenson, A. C.,


Stoneman, R. and Thompson, R. 2000. Replicated
proxy-climate signals over the last 2000 yr from two
distant UK peat bogs: new evidence for regional palaeoclimate teleconnections. Quaternary Science Reviews
19, 481-487.
Bard, E., Raisbeck, G., Yiou, F. and Jouzel, J. 2000. Solar
irradiance during the last 1200 years based on cosmogenic nuclides. Tellus B 52, 985-992.

Bahnson, H. 1968. Kolorimetriske bestemmelser af humieringstal i hjmosetrv fra Fugls Mose p Djursland.
Meddelelser fra Dansk Geologisk Forening 18, 55-63.

Barnekow, L. 2000. Holocene regional and local vegetation history and lake-level changes in the Tornetrsk
area, northern Sweden. Journal of Paleolimnology 23,
399-420.

Baker, A., Caseldine, C. J., Gilmour, M. A., Charman, D.,


Proctor, C. J., Hawkesworth, C. J. and Phillips, N. 1999.
Stalagmite luminescence and peat humication records
of palaeomoisture for the last 2500 years. Earth and
Planetary Science Letters 165, 157-162.

Beer, J., Muscheler, R., Wagner, G., Laj, C., Kissel, C.,
Kubik, P. W. and Synal, H.-A. 2002. Cosmogenic
nuclides during Isotope Stages 2 and 3. Quaternary Science Reviews 21, 1129-1139.

Barber, K., Dumayne-Peaty, L., Hughes, P., Mauquoy,


D. and Scaife, R. 1998. Replicability and variability of
the recent macrofossil and proxy-climate record from
raised bogs: eld stratigraphy and macrofossil data from
Bolton Fell Moss and Walton Moss, Cumbria, England.
Journal of Quaternary Science 13, 515-528.
Barber, K. E. 1981. Peat stratigraphy and climatic change.
A palaeoecological test of the theory of cyclic peat bog
regeneration. A. A. Balkema Publishers; Rotterdam; 219
pp.
Barber, K. E. 1982. Peat-bog stratigraphy as a proxy climate record. In Climatic change in later Prehistory. A. F.
Harding (Ed.). University Press, Edinburgh. 103-113.
Barber, K. E., Battarbee, R. W., Brooks, S. J., Eglinton,
G., Haworth, E. Y., Oldeld, F., Stevenson, A. C.,
Thompson, R., Appleby, P. G., Austin, W. E. N., Cameron, N. G., Ficken, K. J., Golding, P., Harkness, D. D.,
Holmes, J. A., Hutchinson, R., Lishman, J. P., Maddy,
D., Pinder, L. C. V., Rose, N. L. and Stoneman, R. E.

24

Belyea, L. R. and Clymo, R. S. 2001. Feedback control


of the rate of peat formation. Proceedings of the Royal
Society of London: Biological Sciences 268, 13151321.
Berglund, B. E. 1986. Handbook of Holocene Palaeoecology and Palaeohydrology. John Wiley & Sons Ltd.
869 pp.
Bergman, J. 2005. Tree-limit ecotonal response to
Holocene climate Change in the Scandes Mountains of
west-central Sweden. Lundqua Thesis Lund University;
Lund; 35 pp.
Bergman, J., Wastegrd, S., Hammarlund, D., Wohlfarth,
B. and Roberts, S. J. 2004. Holocene tephra horizons at
Klocka Bog, west-central Sweden: aspects of reproducibility in subarctic peat deposits. Journal of Quaternary
Science 19, 241-249.
Bianchi, G. G. and McCave, I. N. 1999. Holocene periodicity in North Atlantic climate and deep-ocean ow
south of Iceland. Nature 397, 515-517.

The colour of climate: changes in peat decomposition as a proxy for climate change

Bjrck, S. and Clemmensen, L. B. 2004. Aeolian sediment


in raised bog deposits, Halland, SW Sweden: a new
proxy record of Holocene winter storminess variation in
southern Scandinavia? Holocene 14, 677-688.

Borgmark, A. and Schoning, K. in press. A comparative


study of peat proxies from two eastern central Swedish
bogs and their relation to meteorological data. Journal
of Quaternary Science

Bjrck, S., Muscheler, R., Kromer, B., Andresen, C. S.,


Heinemeier, J., Johnsen, S. J., Conley, D., Ko, N.,
Spurk, M. and Veski, S. 2001. High-resolution analyses
of an early Holocene climate event may imply decreased
solar forcing as an important climate trigger. Geology
29, 1107-1110.

Boygle, J. 1998. A little goes a long way: discovery of a


new mid-Holocene tephra in Sweden. Boreas 27, 195199.

Blaauw, M., van Geel, B. and van der Plicht, J. 2004. Solar
forcing of climatic change during the mid-Holocene:
indications from raised bogs in The Netherlands.
The Holocene 14, 35-43.
Blackford, J. J. 1998. Holocene climatic Variability in the
North Atlantic Region as shown by Peat Bog Records.
Frskaparrit 46, 155-162.
Blackford, J. J. and Chambers, F. M. 1993. Determining
the degree of peat decomposition for peat-based palaeoclimatic studies. International Peat Journal 5, 7-24.
Blundell, A. and Barber, K. 2005. A 2800-year palaeoclimatic record from Tore Hill Moss, Strathspey,
Scotland: the need for a multi-proxy approach to peatbased climate reconstructions. Quaternary Science Reviews 24, 1261-1277.
Blytt, A. 1876. Essay on the immigration of the
Norwegian ora during alternating rainy and dry periods. Cammermeye; Christiania; 89 pp.
Bond, G., Kromer, B., Beer, J., Muscheler, R., Evans, M.
N., Showers, W., Hoffmann, S., Lotti-Bond, R., Hajdas,
I. and Bonani, G. 2001. Persistent Solar Inuence on
North Atlantic Climate During the Holocene. Science
294, 2130-2136.
Bond, G., Showers, W., Cheseby, M., Lotti, R., Almasi,
P., deMenocal, P., Priore, P., Cullen, H., Hajdas, I. and
Bonani, G. 1997. A Pervasive Millennial-Scale Cycle in
North Atlantic and Glacial Climates. Science 278, 12571266.
Booth, R. K. 2002. Testate amoebae as paleoindicators of
surface-moisture changes on Michigan peatlands: modern ecology and hydrological calibration. Journal of
Paleolimnology 28, 329-348.
Borgmark, A. 2005. Holocene climate variability and
periodicities in south-central Sweden, as interpreted
from peat humication analysis. The Holocene 15, 387395.

Boygle, J. 2004. Towards a Holocene tephrochronology for


Sweden: geochemistry and correlation with the North
Atlantic tephra stratigraphy. Journal of Quaternary
Science 19, 103-109.
Bradley, R. S. 1999. Paleoclimatology: Reconstructing
Climates of the Quaternary. International Geophysics
series 64 Academic Press; 613 pp.
Bradley, R. S., Alverson, K. D. and Pedersen, T. F. 2003a.
Challenges of a Canging Earth: Past Perspectives, Future
Concerns. In Paleoclimate, global change and the future.
K. D. Alverson, R. S. Bradley and T. F. Pedersen (Eds.),
Global change; the IGBP series. Springer Verlag, Berlin.
162-167.
Bradley, R. S., Briffa, K. R., Cole, J., Hughes, M. K. and
Osborn, T. J. 2003b. The climate of the last millennium.
In Paleoclimate, global change and the future. K. D.
Alverson, R. S. Bradley and T. F. Pedersen (Eds.), Global
change; the IGBP series. Springer Verlag, Berlin. 105162.
Briffa, K. R. 2000. Annual climate variability in the
Holocene: interpreting the message of ancient trees.
Quaternary Science Reviews 19, 87-105.
Bronk Ramsey, C. 1995. Radiocarbon Calibration
and Analysis of Stratigraphy: The OxCal Program.
Radiocarbon 37, 425-430.
Burroughs, W., J. 2001. Climate Change. Cambridge University Press; Cambridge; 298 pp.
Carslaw, K. S., Harrison, R. G. and Kirby, J. 2002. Cosmic
rays, clouds and climate. Science 298, 1732-1737.
Caseldine, C., Hatton, J., Huber, U., Chiverrell, R. and
Woolley, N. 1998. Assessing the impact of volcanic activity on mid-Holocene climate in Ireland: the need for
replicate data. The Holocene 8, 105-111.
Caseldine, C. J., Baker, A., Charman, D. J. and Hendon,
D. 2000. A comparative study of optical properties of
NaOH peat extracts: implications for humication studies. The Holocene 10, 649-658.

25

Anders Borgmark

Chambers, F. M., Barber, K. E., Maddy, D. and Brew, J.


1997. A 5500-year proxy-climate and vegetation record
from blanket mire at Talla Moss, Borders, Scotland. The
Holocene 7, 391-399.

Clymo, R. S. 1983. Peat. In Mires: Swamp, Bog, Fen and


Moor. General Studies. Ecosystems of the World 4A:
General Studies. A. J. P. Gore (Ed.). Elsevier, Amsterdam.
159-224.

Chambers, F. M. and Blackford, J. J. 2001. Mid- and LateHolocene climatic changes: a test of periodicity and
solar forcing in proxy-climate data from blanket peat
bogs. Journal of Quaternary Science 16, 329-338.

Clymo, R. S. 1984. The limits to peat bog growth. Philosophical Transactions of the Royal Society of London,
Series B: Biological Sciences 303, 605-654.

Chambers, F. M. and Charman, D. J. 2004. Holocene environmental change: contributions from the peatland
archive. The Holocene 14, 1-6.
Chambers, F. M., Ogle, M. I. and Blackford, J. J. 1999.
Palaeoenvironmental evidence for solar forcing of
Holocene climate: Linkages to solar science. Progress in
Physical Geography 23, 181-204.
Charman, D. 2002. Peatlands and Environmental Change.
John Wiley & Sons Ltd.; Chichester; 301 pp.
Charman, D., J. 2001. Biostratigraphic and palaeoenvironmental applications of testate amoebae. Quaternary
Science Reviews 20, 1753-1764.
Charman, D. J., Caseldine, C., Baker, A., Gearey, B., Hatton, J. and Proctor, C. 2001. Paleohydrological Records
from Peat Proles and Speleothems in Sutherland, Northwest Scotland. Quaternary Research 55, 223-234.
Charman, D. J., Hendon, D. and Packman, S. 1999. Multiproxy surface wetness records from replicate cores
on an ombrotrophic mire: implications for Holocene
palaeoclimatic records. Journal of Quaternary Science
14, 451-463.
Charman, D. J., Hendon, D. and Woodland, W. A. 2000.
The identication of testate amoebae (Protozoa: Rhizopoda) in peats. QRA Technical Guide no. 9 Quaternary
Research Association; London; 147 pp.
Charman, D. J. and Warner, B. G. 1997. The ecology of
testate amoebae (Protozoa: Rhizopoda) in oceanic peatlands in Newfoundland, Canada: modelling hydrological relationships for palaeoenvironmental reconstructions. Ecoscience 4, 555-562.

Cronin, T. M. 1999. Principles of Paleoclimatology.


Columbia University Press; New York; 560 pp.
Dahl-Jensen, D., Mosegaard, K., Gundestrup, N., Clow,
G., D., Johnsen, S., J., Hansen, A., W. and Balling, N.
1998. Past Temperatures Directly from the Greenland
Ice Sheet. Science 282, 268.
Davis, J. C. 1986. Statistics and Data Analysis in Geology. Second Edition. John Wiley & Sons; New York;
646 pp.
Dean, W., Anderson, R., Platt Bradbury, J. and Anderson,
D. 2002. A 1500-year record of climatic and environmental change in Elk Lake, Minnesota I: Varve thickness and gray-scale density. Journal of Paleolimnology
27, 287-289.
Dean, W. E. and Schwalb, A. 2000. Holocene environmental and climatic change in the Northern Great Plains as
recorded in the geochemistry of sediments in Pickerel
Lake, South Dakota. Quaternary International 67, 520.
Denton, G. H. and Karln, W. 1973. Holocene climatic
variations: their pattern and possible cause. Quaternary
Research 3, 155-205.
Digerfeldt, G. 1988. Reconstruction and regional correlation of Holocene lake-level uctuations in Lake Bysjn,
South Sweden. Boreas 17, 162-182.
Dugmore, A. 1989. Icelandic volcanic ash in Scotland.
Scottish Geographical Magazine 105, 168-172.
Dugmore, A. J., Larsen, G. and Newton, A. J. 1995. Seven
tephra isochrones in Scotland. The Holocene 5, 257266.

Chiverrell, R. C. 2001. A proxy record of late Holocene


climate change from May Moss, northeast England.
Journal of Quaternary Science 16, 9-29.

Emeis, K.-C. and Dawson, A. G. 2003. Holocene palaeoclimate records over Europe and the North Atlantic. The
Holocene 13, 305-309.

Clilverd, M. A., Clarke, E., Rishbeth, H., Clark, T. D.


G. and Ulich, T. 2003. Solar activity levels in 2100.
Astronomy and Geophysics 44, 5.20-5.22.

Eronen, M., Hyvrinen, H. and Zetterberg, P. 1999.


Holocene humidity changes in northern Finnish Lapland
inferred from lake sediments and submerged Scots pines
dated by tree-rings. The Holocene 9, 569-580.

26

The colour of climate: changes in peat decomposition as a proxy for climate change

Foster, D. R., Wright Jr, H. E., Thelaus, M. and King, G.


A. 1988. Bog development and landform dynamics in
central Sweden and south-eastern Labrador, Canada.
Journal of Ecology 76, 1164-1185.
Frenzel, B. 1983. Mires repositories of climatic information or selfperpetuating ecosystems? In Ecosystems of
the world 4B mires: swamp, bog, fen and moor general
studies. A. J. Gore (Ed.). Elsevier, Amsterdam. 3565.
Fries, M. 1951. Pollenanalytiska vittnesbrd om senkvartr vegetationsutveckling, srskilt skogshistoria, i nordvstra Gtaland. Acta Phytogeographica Suecica 29;
Uppsala; 220 pp.
Frhlich, C. and Lean, J. 2002. Solar irradiance variability
and climate. Astronomische Nachrichten 323, 203-212.
Granlund, E. 1932. De svenska hgmossarnas geologi.
SGU C 373; Stockholm; 193 pp.
Gunnarson, B. E., Borgmark, A. and Wastegrd, S. 2003.
Holocene humidity uctuations in Sweden inferred from
dendrochronology and peat stratigraphy. Boreas 32,
347-360.
Hammarlund, D., Bjrck, S., Buchardt, B., Israelson, C.
and Thomsen, C. T. 2003. Rapid hydrological changes
during the Holocene revealed by stable isotope records
of lacustrine carbonates from Lake Igelsjn, southern
Sweden. Quaternary Science Reviews 22, 353-370.
Holmgren, K., Karln, W., Lauritzen, S. E., Lee-Thorp,
J. A., Partridge, T. C., Piketh, S., Repinski, P., Stevenson, C., Svanered, O. and Tyson, P. D. 1999. A 3000year high-resolution stalagmite-based record of palaeoclimate for northeastern South Africa. The Holocene 9,
295-309.

Jessen, C. A., Rundgren, M., Bjrck, S. and Hammarlund,


D. 2005. Abrupt climatic changes and an unstable transition into a late Holocene Thermal Decline: a multiproxy lacustrine record from southern Sweden. Journal
of Quaternary Science 20, 349-362.
Karln, W. and Kuylenstierna, J. 1996. On the solar forcing on Holocene climate: evidence from Scandinavia.
The Holocene 6, 359-365.
Karln, W. and Matthews, J. A. 1992. Reconstructing
Holocene Glacier Variations from Glacial Lake Sediments: Studies from Nordvestlandet and JostedalsbreenJotunheimen, Southern Norway. Geograska Annaler
74, 327-348.
Kilian, M. R., van der Plicht, J. and van Geel, B. 1995. Dating raised bogs: New aspects of AMS 14C wiggle matching, a reservoir effect and climatic change. Quaternary
Science Reviews 114, 959-966.
Korhola, A. 1992. Mire induction, ecosystem dynamics
and lateral extention on raised bogs in the southern
coastal area of Finland. Fennia 170, 25-94.
Korhola, A. 1995. Holocene climatic variations in southern Finland reconstructed from peat-initiation data.
The Holocene 5, 43-58.
Kunzendorf, H. and Larsen, B. 2002. A 200-300 year
cyclicity in sediment deposition in the Gotland Basin,
Baltic Sea, as deduced from geochemical evidence.
Applied Geochemistry 17, 29-38.
Langdon, P. G. and Barber, K. E. 2004. Snapshots in time:
precise correlations of peat-based proxy climate records
in Scotland using mid-Holocene tephras. The Holocene
14, 21-33.

Hughes, P. D. M., Mauquoy, D., Barber, K. E. and


Langdon, P. G. 2000. Mire-development pathways and
palaeoclimatic records from a full Holocene peat archive
at Walton Moss Cumbria, England. The Holocene 10,
465-479.

Langdon, P. G., Barber, K. E. and Hughes, P. D. M. 2003. A


7500-year peat-based palaeoclimatic reconstruction and
evidence for an 1100-year cyclicity in bog surface wetness from Temple Hill Moss, Pentland Hills, southeast
Scotland. Quaternary Science Reviews 22, 259-274.

Ingram, H. A. P. 1978. Soil layers in mires: function and


terminology. Journal of Soil Sciences 29, 224-227.

Larocque, I. and Hall, R. I. 2004. Holocene temperature


estimates and chironomid community composition in
the Abisko Valley, northern Sweden. Quaternary Science
Reviews 23, 2453.

Ingram, H. A. P. 1982. Size and shape in raised mire ecosystems: a geophysical model. Nature 297, 300-303.
Ingram, H. A. P. 1983. Hydrology. In Mires: Swamp,
Bog, Fen and Moor. General Studies. Ecosystems of the
World 4A: General Studies. A. J. P. Gore (Ed.). Elsevier,
Amsterdam. 67-158.

Larsen, G., Dugmore, A. and Newton, A. J. 1999. Geochemistry of historical-age silicic tephras in Iceland.
The Holocene 9, 463-471.

27

Anders Borgmark

Larsen, G., Newton, A. J., Dugmore, A. J. and


Vilmundardttir, E. G. 2001. Geochemistry, dispersal,
volumes and chronology of Holocene silicic tephra layers from the Katla volcanic system, Iceland. Journal of
Quaternary Science 16, 119-132.

Marshall, J., Kushnir, Y., Battisti, D., Chang, P., Czaja, A.,
Dickson, R., Hurrell, J., McCartney, M., Saravanan, R.
and Visbeck, M. 2001. North Atlantic climate variability: phenomena, impacts and mechanisms. International
Journal of Climatology 21, 1863-1898.

Lauritzen, S. E. and Lundberg, J. 1999a. Calibration of


the speleothem delta function: an absolute temperature record for the Holocene in northern Norway.
The Holocene 9, 659-669.

Matthews, J. A., Dahl, S. O., Nesje, A., Berrisford, M. S.


and Andersson, C. 2000. Holocene glacier variations in
central Jotunheimen, southern Norway based on distal
glaciolacustrine sediment cores. Quaternary Science Reviews 19, 1625-1647.

Lauritzen, S. E. and Lundberg, J. 1999b. Speleothems and


climate: a special issue of The Holocene. The Holocene
9, 643-647.
Lindeberg, G. 2002. The Swedish varved clays revisited:
Spectral- and Image analysis of different types of varved
series from the Baltic Basin. The Department of Physical
Geography and Quaternary Geology Stockholm University; Stockholm. Thesis. 25 pp.
Lowe, J. J. and Walker, M. J. C. 1997. Reconstructing
Quaternary Environments. Longman; 446 pp.
Lundqvist, G. 1932. Tidvattnet och frsumpningsetapperna. Geologiska Freningens i Stockholm
Frhandlingar 54, 305-309.
Lundqvist, G. 1962. Geological radiocarbon datings from
the Stockholm station. SGU C 589; Stockholm; 23 pp.
Lundqvist, J. 1957. C14-dateringar av rekurrensytor i
Vrmland. SGU C 554; Stockholm; 22 pp.
Lundqvist, J. 1958a. Beskrivning till jordartskarta ver
Vrmlands ln. SGU Ca 38; Stockholm; 229 pp.
Lundqvist, J. 1958b. Studies of the Quaternary history
and deposits of Vrmland, Sweden. SGU C 559; Stockholm; 57 pp.
Lundqvist, J. 1994. Inlandsisens avsmltning. In Sveriges Nationalatlas, Berg och Jord. C. Fredn (Ed.). Bokfrlaget Bra Bcker, Hgans. 124-135.
Magny, M. 2004. Holocene climate variability as reected
by mid-European lake-level uctuations and its probable
impact on prehistoric human settlements. Quaternary
International 113, 65-79.
Mangerud, J., Andersen, S. T., Berglund, B. E. and Donner, J. J. 1974. Quaternary stratigraphy of Norden, a
proposal for terminology and classication. Boreas 3,
109-128.

28

Mauquoy, D. and Barber, K. 1999. A replicated 3000 yr


proxy-climate record from Coom Rigg Moss and Felecia
Moss, the Border Mires, northern England. Journal of
Quaternary Science 14, 263-275.
Mauquoy, D. and Barber, K. E. 2002. Testing the sensitivity of the palaeoclimatic signal from ombrotrophic
peat bogs in northern England and the Scottish Borders.
Review of Palaeobotany and Palynology 119, 219-240.
Mauquoy, D., Blaauw, M., van Geel, B., Borromei, A.,
Quattrocchio, M., Chambers, F. M. and Possnert, G.
2004a. Late Holocene climatic changes in Tierra del
Fuego based on multiproxy analyses of peat deposits.
Quaternary Research 61, 148-158.
Mauquoy, D., Engelkes, T., Groot, M. H. M., Markesteijn,
F., Oudejans, M. G., van der Plicht, J. and van Geel, B.
2002. High-resolution records of late-Holocene climate
change and carbon accumulation in two north-west
European ombrotrophic peat bogs. Palaeogeography,
Palaeoclimatology, Palaeoecology 186, 275-310.
Mauquoy, D., van Geel, B., Blaauw, M., Speranza, A. and
van der Plicht, J. 2004b. Changes in solar activity and
Holocene climatic shifts derived from 14C wiggle-match
dated peat deposits. The Holocene 14, 45-51.
Mayewski, P. A., Meeker, L. D., Twickler, M. S., Whitlow, S., Qinzhao-Yang, Berry-Lyons, W. and Prentice,
M. 1997. Major features and forcing of high-latitude
Northern Hemisphere atmospheric circulation using
a 110 000-year-long glaciochemical series. Journal of
Geophysical Research 102, 26345-26366.
Mayewski, P. A., Rohling, E. E., Stager, C. J., Karln,
W., Maasch, K. A., Meeker, D. L., Meyerson, E. A.,
Gasse, F., van Kreveld, S., Holmgren, K., Lee-Thorpe,
J., Rosqvist, G., Rack, F., Staubwasser, M., Schneider,
R. R. and Steig, E. J. 2004. Holocene climate variability.
Quaternary Research 62, 243-255.

The colour of climate: changes in peat decomposition as a proxy for climate change

Mitchell, E. A. D., van der Knaap, W. O., van Leeuwen, J.


F. N., Buttler, A., Warner, B. G. and Gobat, J. M. 2001.
The palaeoecological history of the Praz-Rodet bog
(Swiss Jura) based on pollen, plant macrofossils and testate amoebae (Protozoa). The Holocene 11, 65-80.
Moberg, A., Sonechkin, D. M., Holmgren, K., Datsenko,
N. M. and Karln, W. 2005. Highly variable Northern
Hemisphere temperatures reconstructed from low- and
high-resolution proxy data. Nature 433, 613-617.

Pilcher, J. R., Hall, V. A. and McCormac, F. G. 1995. Dates


of Holocene Icelandic volcanic eruptions from tephra
layers in Irish peats. The Holocene 5, 103-110.
Reid, G. C. 1987. Inuence of solar variability on global
sea surface temperatures. Nature 329, 142-143.
Romanov, V. V. 1968. Hydrophysics of Bogs. Israel Program for Scientic Translations; Jerusalem; 299 pp.

Mller, H. and Stlhs, G. 1964. Beskrivning till geologiska kartbladet Stockholm NO. SGU Ae 1; 148 pp.

Roos-Barraclough, F., van der Knaap, W. O., van Leeuwen,


J. F. N. and Shotyk, W. 2004. A Late-glacial and Holocene
record of climatic change from a Swiss peat humication
prole. The Holocene 14, 7-19.

Nesje, A., Matthews, J. A., Dahl, O., Berrisford, M. S.


and Andersson, C. 2001. Holocene glacier uctuations
of Flatebreen and winter-precipitation changes in the
Jostedalsbreen region, western Norway, based on glaciolacustrine sediment records. The Holocene 11, 267280.

Rosqvist, G., Jonsson, C., Yam, R., Karln, W. and Shemesh,


A. 2004. Diatom oxygen isotopes in pro-glacial lake
sediments from northern Sweden: a 5000 year record
of atmospheric circulation. Quaternary Science Reviews
23, 851-859.

Nilssen, E. and Vorren, K.-D. 1991. Peat humication and


climate history. Norsk Geologisk Tidsskrift 71, 215217.

Sandegren, R. and Magnusson, N. H. 1937. Beskrivning


till kartbladet Forshaga. SGU Aa 179; 117 pp.

Nilsson, T. 1964. Entwicklungsgeschichtliche Studien im


Agerds Mosse, Schonen. Lunds universitets rsskrift.
Andra avdelningen, Medicin samt matematiska och
naturvetenskapliga mnen 59; Lund; 34 pp.

Schoning, K., Charman, D. J. and Wastegrd, S. 2005.


Reconstructed water tables from two ombrotrophic
mires in eastern central Sweden compared with instrumental meteorological data. The Holocene 15, 111118.

OBrien, S. R., Mayewski, P. A., Meeker, L. D., Meese, D.


A., Twickler, M. S. and Whitlow, S. I. 1995. Complexity
of Holocene climate as reconstructed from a Greenland
ice core. Science 270, 1962-1964.

Schulz, M. and Mudelsee, M. 2002. REDFIT: estimating


red-noise spectra directly from unevenly spaced paleoclimatic time series. Computers & Geosciences 28, 421426.

Oldeld, F. 2004. Some personal perspectives on the role


of Quaternary science in global change studies. Global
and Planetary Change 40, 3-9.

Schulz, M. and Stattegger, K. 1997. SPECTRUM: Spectral


analysis of unevenly spaced paleoclimatic time series.
Computers & Geosciences 23, 929-945.

Oldeld, F., Thompson, R., Crooks, P. R. J., Gedye, S. J.,


Hall, V. A., Harkness, D. D., Housley, R. A., McCormac,
F. G., Newton, A. J., Pilcher, J. R., Renberg, I. and Richardson, N. 1997. Radiocarbon dating of a recent highlatitude peat prole: Stor myrn, northern Sweden.
The Holocene 7, 283-290.

Sepp, H., Nyman, M., Korhola, A. and Weckstrm, J.


2002. Changes of the tree-lines and alpine vegetation
in relation to post-glacial climate dynamics in northern
Fennoscandia based on pollen and chironimid records.
Journal of Quaternary Science 17, 287-301.

Osvald, H. 1923. Die Vegetation des Hochmoores Komosse. Svenska Vxtsociologiska Sllskapets Handlingar 1,
1-436.

Sernander, R. 1908. On the evidence of postglacial


changes of climate furnished by the peat-mosses of
northern Europe. Geologiska Freningens i Stockholm
Frhandlingar 30, 467-478.

Persson, C. 1966. Frsk till tefrokronologisk datering


av ngra svenska torvmossar. Geologiska Freningens i
Stockholm Frhandlingar 88, 361-394.

Sernander, R. 1909. De scanodaniska torfmossarnas stratigra. Geologiska Freningens i Stockholm Frhandlingar


31, 423-448.

Pster, C., Schwarz-Zanetti, G., Wegmann, M. and Luterbacher, J. 1998. Winter air temperature variations in
western Europe during the Early and High Middle Ages
(AD 750-1300). The Holocene 8, 535-552.

Sernander, R. 1912. Ausstellung zur Beleuchtung der Entwicklungsgeschichte der schwedischen Torfmoore. In
XI Congrs Gologique International. Kungl. Boktryckeriet. P. A. Norstedt & Sner, Stockholm 1910. 1413
pp.

29

Anders Borgmark

Snowball, I., Korhola, A., Briffa, K. R. and Ko, N. 2004.


Holocene climate dynamics in Fennoscandia and the
North Atlantic. In Past Climate Variability through Europe and Africa. R. W. Battarbee (Ed.). Kluwer Academic
Publishers, Dordrecht. 465-494.
Snowball, I., Sandgren, P. and Petterson, G. 1999. The
mineral magnetic properties of an annually laminated
Holocene lake-sediment sequence in northern Sweden.
The Holocene 9, 353-362.
Stuiver, M. and Braziunas, T. F. 1989. Atmospheric 14C and
century-scale solar oscillations. Nature 338, 405-408.
Stuiver, M. and Braziunas, T. F. 1998. Anthropogenic
and solar components of hemisperic 14C. Geophysical
Research Letters 25, 329-332.
Stuiver, M., Reimer, P. J., Bard, E., Beck, J. W., Burr, G.
S., Hughen, K. A., Kromer, B., McCormac, G., van
der Plicht, J. and Spurk, M. 1998a. INTCAL98 Radiocarbon Age Calibration, 24000-0 cal BP. Radiocarbon
40, 1041-1083.
Stuiver, M., Reimer, P. J. and Braziunas, T. F. 1998b. High
precision radiocarbon age calibration for terrestial and
marine samples. Radiocarbon 40, 1127-1151.
Svantesson, S.-I. 1981. Beskrivning till jordartskartan Hjo
SO. SSGU Ae Nr 44; 101 pp.
Svensson, G. 1988. Bog development and environmental
conditions as shown by the stratigraphy of Store Mosse
mire in southern Sweden. Boreas 17, 89-111.
Telford, R. J., Heegaard, E. and Birks, H. J. B. 2004. All
age-depth models are wrong: but how badly? Quaternary
Science Reviews 23, 1-5.
Thelaus, M. 1989. Late Quaternary vegetation history and
palaeohydrology of the Sandsjn-rshult area, southwestern Sweden. Lund University. Thesis. 76 pp.
Tipping, R. 1995. Holocene evolution of a lowland
Scottish landscape: Part I, peat and pollen-stratigraphic evidence for raised moss development and climatic
change. The Holocene 5, 69-81.
Tolonen, K. 1973. On the rate and pattern of peat formation during the Postglacial time. Suo 24, 83-88.
Tolonen, K. 1987. Natural history of raised bogs and forest vegetation in the Lammi area, southern Finland,
studied by stratigraphical methods. Annales Academiae
Scientiarum Fennicae A III 144, 1-46.

30

Tolonen, K., Huttunen, P. and Jungner, H. 1985. Regeneration of two costal raised bogs in eastern North
America. Annales Academiae Scientiarum Fennicae A
III 139, 1-51.
van den Bogaard, C., Drer, W., Glos, R., Nadeau, M.,
Grootes, P. and Erlenkeuser, H. 2002. Two tephra layers
bracketing Late Holocene palaeoecological changes in
northern Germany. Quaternary Research 57, 314-324.
van den Bogaard, C. and Schmincke, H.-U. 2002. Linking
the North Atlantic to central Europe: a high-resolution
Holocene tephrochronological record from northern
Germany. Journal of Quaternary Science 17, 3-20.
van Geel, B., Heusser, C. J., Renssen, H. and Schuurmans,
C. J. E. 2000. Climatic change in Chile at around 2700
BP and global evidence for solar forcing: a hypothesis.
The Holocene 10, 659-664.
van Geel, B., Raspopov, O. M., Renssen, H., van der
Plicht, J., Dergachev, V. A. and Meijer, H. A. J. 1999.
The role of solar forcing upon climate change. Quaternary Science Reviews 18, 331-338.
van Geel, B., van der Plicht, J., Kilian, M. R., Klaver, E.
R., Kouwenberg, J. H. M., Renssen, H., Reynaud Farrera, I. and Waterbolk, H. T. 1998. The sharp rise of
14C ca. 800 cal BC: Possible causes, related climatic teleconnections and the impact on human environments.
Radiocarbon 40, 535-550.
Warner, B. G. and Charman, D. J. 1994. Holocene changes
on a peatland in northwestern Ontario interpreted from
testate amoebae (Protozoa) analysis. Boreas 23, 270279.
Warner, B. G., Clymo, R. S. and Tolonen, K. 1993. Implications of Peat Accumulation at Point Escuminac, New
Brunswick. Quaternary Research 39, 245-248.
Wastegrd, S. 1998. Deglaciation chronology and marine
environments in southwestern Sweden. Boreas 27, 178194.
Wastegrd, S. 2002. Early to middle Holocene silicic
tephra horizons from the Katla volcanic system, Iceland:
new results from the Faroe Islands. Journal of Quaternary Science 17, 723-730.
Wastegrd, S. 2005. Late Quaternary tephrochronology
of Sweden: a review. Quaternary International 130, 4962.

The colour of climate: changes in peat decomposition as a proxy for climate change

Wastegrd, S., Bjrck, S., Grauert, M. and Hannon, G.


E. 2001. The Mjuvtn tephra and other Holocene tephra horizons from the Faroe Islands: a link between the
Icelandic source region, the Nordic Seas, and the European continent. The Holocene 11, 101-109.
Weber, C. A. 1926. Grenzhorizont und Klimaschwankungen. Abhandlungen Naturwissenschaftlicher Verein
Bremen 26, 98-106.
Vedin, H. 2004. Lufttemperatur. In Klimat, Sjar och
Vattendrag. B. Raab and H. Vedin (Eds.), Sveriges
Nationalatlas. Bokfrlaget Bra Bcker, Hgans. 4457.
Viau, A. E., Gajewski, K., Fines, P., Atkinson, D. E. and
Sawada, M. C. 2002. Widespread evidence of 1500 yr
climate variability in North America during the past
14000 yr. Geology 30, 459462.
Wijmstra, T. A., Hoekstra, S., de Vries, B. J. and van der
Hammen, T. 1984. A preliminary study of periodicities
in percentage curves dated by pollen density. Acta Botanica Neerlandica 33, 547-557.
von Post, L. 1921. Upplysningar rrande Sveriges geologiska undersknings torvmarkrekognoscering. Beskrivning till kartbladet Upperud, SGU D 52; I-XVI.
von Post, L. and Granlund, E. 1926. Sdra Sveriges
Torvtillgngar I. SGU C. 335; Stockholm; 127 pp.
von Post, L. and Sernander, R. 1910. Panzenphysiognomische Studien aus Torfmooren in Nrke.
XI International Geological Congress: Excursion Guide
No. 14 (A7); Stockholm; 48 pp.
Woodland, W. A., Charman, D. J. and Sims, P. C. 1998.
Quantitative estimates of water tables and soil moisture in Holocene peatlands from testate amoebae.
The Holocene 8, 261-273.
Yu, S.-Y. 2003. Centennial-scale cycles in middle Holocene
sea level along the southeastern Swedish Baltic coast.
Geological Society of America Bulletin 115, 1404
1409.
Zilln, L., Wastegrd, S. and Snowball, I. 2002. Calendar
year ages of three mid-Holocene tephra layers identied in varved lake sediments in west central Sweden.
Quaternary Science Reviews 21, 1583-1591.

31

Вам также может понравиться