Вы находитесь на странице: 1из 172

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/236024084

WATER POLLUTION - methods and criteria to rank, model and remediate


chemical threats to aquatic ecosystems

Book · January 2008

CITATIONS READS

93 1,214

2 authors:

Lars Håkanson Andreas Bryhn


Uppsala University Swedish University of Agricultural Sciences
475 PUBLICATIONS   11,450 CITATIONS    97 PUBLICATIONS   649 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Nutrient dynamics in the Baltic Sea. View project

The AWARE project View project

All content following this page was uploaded by Lars Håkanson on 31 May 2014.

The user has requested enhancement of the downloaded file.


WATER POLLUTION

- methods and criteria to rank, model and remediate chemical


threats to aquatic ecosystems

Lars Håkanson and Andreas Bryhn

Part 1. The PER-approach - criteria to rank threats to aquatic


ecosystems
Part 2. Introduction to aquatic ELS-modeling

Uppsala University Printed at


Department of Earth Sciences Geotryckeriet
SWEDEN Uppsala, 2008
 
Contents
Prologue iv
1. The PER-approach - criteria to rank threats to aquatic ecosystems 1
1.1. Introduction and aim 1
1.1.1. Background on chemical threats, especially metals 6
1.1.2. The PER-concept 9
1.2. Ecosystem analyses - basic concepts 12
1.2.1. Defining ecosystem boundaries 12
1.2.2. Ecosystem indices 14
1.2.3. Environmental threats 15
1.2.4. Target ecosystems 16
1.2.5. Regression models and mass-balance models 21
1.3. Acidifying substances 23
1.3.1. Background 23
1.3.2. The geographical perspective 24
1.3.3. Acidification and metals 28
1.3.4. The time perspective 30
1.3.5. Summary - acidification 32
1.4. Metals, 32
1.4.1. Brief background on metals 32
1.4.1. Mercury 33
1.4.1.1. Effects variables for mercury in aquatic ecosystems 33
1.4.1.2. Geographical perspective 34
1.4.1.3. Effect-load-sensitivity 35
1.4.1.3.1. Practical use of ELS models in contexts of remediation 38
1.4.1.4. Time perspective 42
1.4.1.5. Cause and effect - a chemical theory 45
1.4.1.6. Remedial measures 48
1.4.1.7. Summary - mercury 50
1.4.2. Radiocesium 50
1.4.2.1. Effect variables 50
1.4.2.2. Geographical perspective 54
1.4.2.3. Effect-load-sensitivity 54
1.4.2.4. Time perspective 56
1.4.2.5. Summary - radiocesium 58
1.5. Chlorinated organics 59
1.5.1. Background 59
1.5.2. Effect-load-sensitivity 62
1.5.3. Geographical perspective 66
1.5.4. Time perspective 69
1.5.5. Summary - chlorinated organics 71
1.6. Nutrients 71
1.6.1. Introduction to eutrophication 71
1.6.2. Marine eutrophication 73
1.6.2.1. Effect-load-sensitivity 79

ii
1.6.2.2. The geographical perspective 84
1.6.2.3. The time scale 85
1.6.3. Summary - coastal eutrophication 85
1.6.4. Lake eutrophication 86
1.6.4.1. Effect-load-sensitivity 86
1.6.4.2. Area and time perspectives 90
1.6.4.3. Summary - lake eutrophication 90
1.7. Conclusions - a ranking of the threats 92
2. Introduction to aquatic ELS modeling 95
2.1. Background and aim 95
2.1.1. The role of prediction 95
2.1.2. An unambiguous definition of scientific method 95
2.1.3. Testable predictive models 96
2.2. Ecosystem sensitivity 102
2.2.1. Basic hydrodynamic principles and processes for coastal areas 102
2.2.2. Fundamental sedimentological principles and processes for coastal areas 105
2.2.3. A coastal sensitivity index (SI) 108
2.2.4. Lake sensitivity 110
2.3. Time and area compatibility of data 110
2.4. Statistical aspects of regression analysis 115
2.4.1. The ecometric matrix 115
2.4.2. Confidence intervals and frequency distributions 117
2.4.3. Prairie's "staircase" 119
2.4.4. Other statistical concepts and aspects 119
2.5. Variability and uncertainty 122
2.5.1. Variability within and among aquatic ecosystems 122
2.5.2. The sampling formula and uncertainties in empirical data 124
2.6. Principles determining the predictive success of ecosystem models 126
2.6.1. The highest possible r2 from Emp1-Emp2; re2 127
2.6.2. Highest reference r2; rr2 131
2.6.3. Comparing model predictions with re2 and rr2 132
2.7. Dynamic and static ecosystem modeling 133
2.7.1. The classical ELS model - lake eutrophication 133
2.7.2. ELS modeling of coastal eutrophication 137
2.8. Model testing 149
2.8.1. Calibration and validation 149
2.8.2. Sensitivity tests 152
2.8.3. Uncertainty tests using Monte Carlo techniques 154
2.8.4. Uncertainty and sensitivity analysis as tools for structuring ELS models 161
3. Epilogue 164
Literature references 165

Reaching the goal is certainly important for the ride


but it is the travelling itself that gives content and pleasure to the stride.
(after Karin Boye)

iii
Preface to the edition of 2008

Humans have altered the aquatic environment in a large number of ways, and we regularly get exposed
to a multitude of information about various environmental problems via mass media and other sources.
These problems vary in terms of severity, geographical spread and duration. Small problems can
indeed be solved by individuals, while more complex and widespread problems need large-scale
abatement strategies from communities of various sizes and extents. Environmental management
concerns the latter type of problems and a basic purpose of environmental management is to direct
time and effort towards large environmental problems rather than small and/or imaginary ones.

There are two old proverbs that wittingly illustrate which strategies we should avoid if we as
professionals really want to achieve something substantial in practice and make a change; one of
these proverbs is "not seeing the forest for the trees", and the other one is "straining out gnats
while swallowing camels". This book aims at defining methods for pointing out the "camels" among
problems in the aquatic environment and at providing tools to prevent, or at least decrease, the
swallowing of these camels.

To do this, it is crucial to have a system of criteria to structure, analyze and dimension the problems.
Subjective approaches are inadequate to the challenges of environmental management. Fully objective
scientific methods and results are generally only applicable for specific substances in specific contexts.

This book focuses not on the conditions at specific sites but at the ecosystem level, on effect-load-
sensitivity (ELS) analysis, on geographical patterns (distribution over area and time), and on
practically useful models to aquatic ecosystem management The examples concern the major threats
to aquatic ecosystems, such as acidification, eutrophication and contamination and the case studies
deal with Swedish lakes and coastal areas. Many of the principles, however, apply to other types of
ecosystems, countries or regions.

It is often argued that the quality of science is related to the possibility to make meaningful
predictions. This argument would give chemistry and physics a pool position in the scientific
community and, for example, economics, a low rating. Robert Peters (Peters, 1991) has convincingly
shown that many aspects of ecology would belong to the same category as economics, or even
theology. But how would predictive ecology rate? This book can be classified as an example of
predictive ecology.

It has long been argued that due to the complex nature of ecosystems, it will never be possible to
predict important target variables, especially not with more comprehensive dynamic models. This
book will demonstrate that those arguments are no longer valid. The key lies in the structuring of
the predictive models. The aim of this book is to discuss just that and to carry out these discussions
within a specific framework where the ultimate aim is to produce meaningful tools for practical water
management

This book is mainly based on the first two chapters in "Water Pollution" by Lars Håkanson (1999).
The second chapter has been revised by Andreas Bryhn and Lars Håkanson in 2008 and now includes
examples from Andreas Bryhn's (2008) PhD thesis as well as from "Tools and Criteria for Sustainable

iv
Coastal Ecosystem Management" by Håkanson and Bryhn (2008) - on behalf of some deleted parts of
the previous edition of this book.

The selection of examples does not mean that the authors disregard the fact that much important
related research has been carried out elsewhere by scientists not cited in this book. The point is instead
that the case-studies exemplified in this book very well fulfill the purpose of efficient ELS analysis
and that data from these examples can be easily distributed by us without permission from other
researchers.

Some readers may fear that some of the information in this book is out of date, since many references
concern conditions as they were during the 1990s. However, the fact is that the roots and extent of
virtually all of the mentioned problems have changed very little since then.

Eutrophication in the Baltic Sea is persistent. Great advances in combating lake eutrophication were
made in the 1970s and 1980s, although much of what remained in the 1990s still remains in 2008 with
respect to this problem. Severe and widespread ecosystem effects from lake acidification are still
routinely documented although sulphur emissions have drastically decreased. The recovery process
from lake acidification is apparently enduring and non-linear. Similarly, high concentrations of
some harmful pollutants are still evident in a very large number of lakes and coastal areas despite
the fact that much effort has been spent on remediating these problems.

The possible exception is radiocesium contamination, a problem which has a low ranking in this book
based on information from the 1990s - and the low ranking remains in 2008. The potential ecological
risk (PER) value should have decreased somewhat compared to the already low value from 1999,
because of the diminishing effects from the Chernobyl accident in 1986, which even at that time
were rather limited in the Swedish aquatic environment compared to effects from other problems.
Thus, the methodology used in this book, alongside the ecosystem processes, the facts on the
ground, and the ranking of chemical threats are still indeed utterly relevant.

Part 1 of this book corresponds to 1 week of full-time studies. The focus is on the PER-system, a
broad, holistic diagnostic system to structure and rank chemical threats to aquatic ecosystems. A
central part of the PER-system concerns effect-load-sensitivity models (ELS), but part 1 does not
discuss mathematical and/or statistical aspects of ELS-models. A basic knowledge on aquatic systems
may, however, be helpful to understanding the text.

Part 2 also corresponds to 1 week of full-time studies. This part concerns the basic elements of ELS-
models, especially dynamic mass-balance models and statistical regression models. Knowledge of
basic statistics and mathematics is needed to fully understand this text. The book thus corresponds to 2
weeks of full-time studies.

v
1. The PER-approach - criteria to rank threats to aquatic
ecosystems
1. 1. Introduction and aim

A few introductory figures will be used to illustrate the basic approach of the flfst part of this tex~ the PER-
system (Potential Ecological Risk) to structure, analyse and rank chemical threats to aquatic ecosystems. The
effect-load-sensitivity analysis (ELS) is an important part of this system. Given a certain threat (load = dose) to
a complex ecosystem, crucial questions are:

• Which are the target effect variables? and Why?


• How can chemical threats to ecosystems be quantitatively ranked?
• How is it possible to develop general, validated, predictive ELS-models for the target effect variables based on
as few, simple and readily available driving variables as possible?
• How can such ELS-models be used as tools to optimise remedial strategies so that the costs for these measures
can be quantitatively related to the environmental benefit?
• How can such Imowledge be accessible to people responsible for environmental management?

All chemical threats cannot be equally important. Subjective criteria to rank are insufficient How is it possible to
develop and apply more objective and scientifically warranted criteria to rank chemical threats to aquatic
ecosystems? That is the key issue addressed in section 1 of this book.

One and the same load of a pollutant may cause very different effects in ecosystems of different sensitivities (=
vulnerabilities). In this book, many examples related to the major chemical threats to aquatic ecosystems, such as
acidification, eUlTophication and toxic contamination, will be given. Crucial questions are: How are effect, load and
sensitivity variables operationally defined? How is the ecosystem defined?

The complicated nature of ecosystems makes it very difficult indeed to carry out causal, mechanistic
analyses concerning the quantitative linkages between a given threat (like a contamination of nutrients)
and variables expressing ecosystem effects. This means that it is very important to define the specific goais
and to apply a slTUctured analysis to reach tilOse goals. This is what this book is all about.

The problems in complex ecosystems may seem insurmountable, and we will use an initial example to il!uSlTate this.
The rest of the book deals with methods to handle such complexities to reach certain defined goals in ranking,
predicting and remediating water pollution.

Fig. 1.1 exemplifies that tile salinity is a most important abiotic sensitivity factor for the biology in coastal areas. The
salinity of the water influences water slTatification, mixing, and hence also the distribution and effects of chemical
pollutants. The salinity may also be regarded as a water chemical "cluster" variable in the sense that there are many
other water chemical variables that are causally or functionally related to the salinity, like the hardness of the water.
Also tile oxygen concenlTation is a most important abiotic variable (see upper right corner of fig. 1.1): The number of
species and individuals of the bottom fauna decrease markedly as the load of organic matter increase, and the oxygen
concenlTation in tile sediments decrease. So. ulere exist a clear and direct relationship between a "simple" abiotic
variable, 02-concenlTation. and lllrget effects. like ule extinction of key functional organisms of Ule bottom fauna.

1
Ecological effect variables

=== ==""' lem}

Increased contounlnation of organic materta.l.3


Decreased oxygCl COncentration
OSlOl:S:l.02S~

SalInJty (%o}

Conductivity Hardness ......._-------~"


Alloilinlty Phytoplankton
Sallnlty Algae
Zoop1a.nk1on
No

Ca CHEMISTRY BIOLOGY
Fish
COASTAL
K
Peripbyton
ECOSYS1EM
Fe Phosphorus Chlorophyll
pH Oxygen Depth Bottom fauna
Coastal_
PHYSICS Saiimentation
Volume Sec:ch1depth
Topograph1cal openness
1~70_1 ~1B~ I
WaleI' exclmoge

Water depth
(m) 3·15_

--.. " .• - n • - "


0-3

.
'0

0
EroSion

r.lD.!!portatlon:
...'"os
. "=un 000
••• • •• .. (-)
so Exn=ely Exn=ely Extremely
Valuable
v:;?;
-
valuable valuable tD
a= valua Ie
area area

Iofauna production

Chemical threat. e.g.. from nutrients in coastal areas ~

Fig. 1.1. Illustration of the complex interactions between various chemical, biological and physical factors that
may be used to characterize a coasial ecosystem ("Everything depends on everything else"). Example:
Top left: The relationship between salinity and number of species. Redrawn from Remane (1934).
Top right: The relationship between load of organic malerial, oxic conditions and benthic communities in a
marine environment. Redrawn from Pearson and Rosenberg (1976).
Lower left: The ETA-diagram (erosion, transportation, accumulation). From Hiikanson and Jansson (1983).
Lower ri"ht A model to estimate the fishery biological value of a given coastal area. Redrawn from Hllkanson
and Rosenberg (1985). Figure from Hiikanson (1991b).

2
Fig. 1.1 (lower, right) shows that the organic content of the sediment, the sediment type, the water depth and the
topographical openness of the coast are linked to the production capacity (the ratio between the production and the
biomass). At large water depths (15-70 m) the production of bottom animals is rather low (= 1). At shallow waters the
production can be very high, especially in semi-enclosed bays with mixed sediments and in estuaries. This depends on
the type of the sediments, Ille habitat for Ille bottom fauna And the sediment type depends largely on the relationship
between the effective fetch and the water depth (see fig. 1.1, lower left diagram). The effective fetch is a measure of the
free open water surface over which the winds may act upon the waves; the larger the effective fetch, the higher the
waves, the larger the wave energy and Ille greater the capacity of the waves to erode and transport the material on the
bottom. It is evident that most of the variables illustrated in fig. 1.1 are more or less related to each other in an
extremely complex web of relationships.

To establish such relationships quantitatively, it is essential to use a rational and structured analysis, and this is
what the following section would like to convey.

Another motive for this book is to highlight problems concerning causal analyses, i.e., the problem in science to
differentiate between cause and effect. Fig. 1.2 offers an example of this concerning an important issue in water
management: Effects of land-use, in Ibis case extensive lumbering, on water quality, in this case lake water
colour. It is evident Illat many activities in the catchment area influence the runoff of substances from land to
water, e.g., coloured substances, like humic materials, metals and nutrients. The aim here is neilller to discuss this
problem nor to exatrtine the processes regulating the transport of water, ions, substances and materials from the
catchment, into rivers and lakes. The aim is simply to give one example of tlle importance of understanding the
concept of cause and effect, and the need for validated predictive models.

Fig. 1.2A illustrates the month-to-month variation in lake colour in lake 2106, Stora Kr6ntjfun. Because the
catchment area of ulis lake was extensively lumbered during late 1986 and early 1987, it is tempting to conclude
Ulat ulOse extensive and easily identifiable operations caused the drastic increase in lake colour observed in fall
1986. That surmise would, however, be more wrong than right. Fig. 1.2B demonstrates that more or less exactly
Ule same seasonal pattern in lake colour could be seen in another lake, lake 2105, Hoimsj6n. There were no
timber operations at all in that catchment area during these years. The increased values of lake colour during
fall, and the general pattern in lake 2105 during 1986, is not linked primarily to forestry but to regional climate
and hydrology. It is likely Illat tree-harvesting increases the transport of coloured substances to lakes. But it is also
a fact Illat such increases cannot be clearly distinguished from other factors in this case. Fig. I.2C shows that lake
colour fluctuated markedly among different years (1986 to 1989) in lake 2106 depending on fluctuations in
precipitation, temperature. etc.

The lesson from this example is that verbal "causal explanations" for certain phenomenon are temptingly easy
when only one or few factors (= facts) are known. In fact, causal explanations of ecosystem level phenomena are
very difficult. The concentrations of a given chemical variable (e.g., lake colour) or the amount of some biological
variable are so frequenUy multiply determined that we should suspect facile explanations. The answer to this
scientific question depends in ecosystem contexts very much on the scale in focus and On Ule availability of
quantitative models which include all the key processes. Had such a model been available, it would be possible to
conduct simulations to see if it is realistic to surmise !bat the land-use operation in this example is likely to be
responsible for Ule increase in lake colour and nOl, e.g., increased runoff from heavy precipitation.

3
Lake 2106. Stofa Krontjiirn
A
........ 180
'a.S-
.§.
'00 140
160
Lumbering

---------~'" ~,
r-- 199&

~-- ~ Effect oflumbenng?


120
S 100
o 80
ao 60
u 40
.I:
n1 20
....l

2 3 4 5 6 7 8 9 10 II 12
Month
200 B.
5- 180t-_-~" Lake 2105. Holmsjon
a. 160
'00 140 ---1986
.§. 120
J- 100
5 80
'8 60
.I: " 40
20
No cutting or ditching
j
0+--+--+--+--+--+--+--r--+-~__~4
2 3 4 5 6 7 8 9 10 II 12
Month
......... 180
'S-
a. 160 ---1986
'00 140 -----1987
.§. 120
5 100
o 80 ------1989
'0o 60
"
.I: 40
'"
....l 20

2 3 4 5 6 7 8 9 10 II 12
Month

Fig. 1.2 A. Seasonal variability in lake colour (based on monthly data) in lake 2106 during 1986. Extensive
cutting operations were carried out in this catchment area during fall 1986 and early 1987.
B. Seasonal variability in lake colour (based on monthly data) in lake 2105 in 1986. No cutting operations were
carried out in this catchment area.
C. Seasonal variability in lake colour (based on monthly data) in lake. 2106, over four years, 1986 to 1989.
Based on Hilkanson and Peters (1995).

There are at least three different approaches based on three different scales to !be issues discussed in this book:

1. The mesocosm-scale concerns artificial but realistic micro worlds (see fig. 1.3).
2. The ecosystem-scale concerns real, natural ecosystems of a given extent in area (0.01 10 1000 km 2 ) and time
(days to years).
3. The multi-media-scale including water, biota, sediments, aunosphere, i.e., large geographical areas or long
periods of time (Mackay, 1979; Mackay and Paterson, 1982).

The focus of this work is at the ecosystem scale, and not at other smaller or larger scales.

4
V
ft
c
,
:.:.: ........... ~ .
r.!...........
.·..........
.............
· .......... .
.. .........
·....... ... ...
..........
........ ... ..
...........
...........
. . . . . . . . . ....
.·............
...........
...........
........... . . . . . ...
. . . ., .......
. . . .....

A =Reservoi r B = Si phon C = Ges trep

Fig. 1.3. An example of a mesocosm. A model of the shallow-water ecosystem of the Baltic in an outdoor basin (8
m 3). The most sensitive key functional organisms for a given chemical contamination are the target organisms.
The concentrations that cause effecLS on the target organisms are the critical concentrations. From Hiikanson
(1990a).

Table 1.1. Various chemical ulreaLS and some examples of ecological effecLS. There are also many physical
threaLS to aquatic ecosystems, like the building of dams, piers and marinas, and many biological threats, sucb
as the intrnduction of new species.

Chemical threat Ecological effects


1. Acidification Increase in fUamentous algae
Reduced reproduction of crustaceans, snails, bivalves and roach

2. Eutrophication Decrease in Secchi depth


Increase in cbloroJlhyll-a and
Hypolimnetic oxygen demand'

3. Contamination
3.1. Metals Increased concentration in fish for human consumption
3.2. Radionuclides Decrease in reproduction of key organisms, e.g .. zooplankton, benthos and fish
3 '1 Qroanic toxins

The effect variable is a key concept in this approach. It is of vital importance that tile reader realises what is
meant and not meant by the effect variable. The effect variable can be identified by mesocosm studies and
then applied in the real world, the ecosystem. A mesocosm is a reproduction of the real ecosystem (e.g., a given
lake type) that is as close to reality as possible in a reasonably large "laboratory scale" (Landner, 1989; Lehtinen
et al .. 1996, 1998). The mesocosm should contain tile fundamental functional groups which form and characterize
the actual ecosystem. In this connection, the purpose of the mesocosm is to study, under controlled conditions, the
substance of interest in order to see (1) which parts of the ecosystem are first damaged and (2) the
concentrations at which the damage occurs. Naturally, it is not possible to simulate in a mesocosm everytiling
tim bappens in nature, such as the influence of weather, wind, currents and other animal species than those
included in the mesocosm. However, it is important that tile mesocosm studies lead to identifying target effect
variables and critical concentrations because this is essential for studies in real, natural ecosystems, the target
scale in practical water management and in tilis book.

5
So, tbis book focuses on tile ecosystem-scale, and on tbe major chemical threats to aquatic ecosystems (table
l.l):
1. Acidification,
2. Eutrophication, and
3, Contamination (of metals, organic toxins and radionuclides).

• In acidification research, tbe target effect variables could be measures of reproduction damage to the most
sensitive fish species (like roach), and target operational variables could be natural or preindustrial average values
of lake pH and/or alkalinity. This information is important in remedial liming because there is no benefit to raising
pH above tbe natural level and because excessive liming increases cost needlessly. The economics of liming are a
major concern in Sweden where the costs for lake liming to control the effects of antbropogenic acidification now
exceed 150 million Swedish SKR (about 20 -25 million U.S. dollars) per year. In tbe U.S.A .. tbe narurallevel was
a target variable, because of the intense debate about whetber cultural acidification had even occurred.

• In eutrophication, likely target variables are mean, representative ecosystem values of chlorophyll-a (a
practical, operational measure of algal biomass and an indicator of primary productivity), or Secchi depth (as a
general index of lake trophy) and oxygen concentration in tbe bottom water. The concentration of phosphorus in
lakes can be used to predict all tbese target variables for lake eutrophication.

• In Ecotoxicology, tbe target variables might include mean ecosystem concentrations of toxic substances in fish
destined for human consumption (such concentrations are often used to set guidelines, blacklisting limits and
environmental goals) and operationally defined, ecological effects on key organisms, like mortality, reproduction
and abundance of important functional groups in defined ecosystems.

• In Radioecology, tbe goal may be to measure and predict tbe concentrations or activities of radionuclides in (1)
water used for irrigation and (2) in fISh for human consumption. The emphasis is not prediction for specific
instances or samples (e.g., for Fish B from location A at time X) but ratber to get a good quantitative picrure for
larger areas over longer periods of time, Le., for entire ecosystems.

1.1.1. Background on chemical threats, especially metals

This book will use mercury and radiocesium as type elements and lakes and coastal areas as type ecosystems to
illustrate many of the fundamental principles and processes regulating tbe spread, biouptake and ecosystem effeclS
of contaminates in general. Mercury belongs to a group of elements often referred to as heavy metals (Le., metals
with a density> 5/cm 3). These metals generally form oxides and sulphides which are often very hard to dissolve,
and Illey tend to be bound in stable complexes witb organic and inorganic particles, Ille "carrier particles". TIle
great interest in heavy metals in aquatic ecotoxicology derives from the fact that some of Illese elements are
supplied to water systems in great excess by man, and tilat some of them are hazardous to the aquatic life because
(see Bowen, 1966; Forstner and Muller, 1974; Forsmer and Witunann, 1979; Salomons and F6rsmer, 1984):

• Tbey can disturb enzymatic systems because of the high electro-negative affinity for reactive groups on ille
enzymes, like amino or sulfhydryl groups.
• They can form stable complexes with essential metabolites.
• They can catalyse ille breakdown of such metabolites.
• They can permeate the membranes of cell.
• Metals can also substitute oiller elements wiill important functions in Ille cell metabolism.

6
Table 1.2. Tbe abundance of various elemenlS (in ppm) in igneous rocks, soils, fresh water, land planlS and
land animals (from Bowen, 1966).

Igneous rocks Soils Fresh water Land plants Land animals

Ag om 0.1 0.00013 0.06 0.006


Al 82,000 71,000 0.24 500 4-100
As 1.8 6.0 0.004 0.2 <; 0.2
Cd 0.2 .0.06 <0.08 0.6 <; 0.5
Co 25 8.0 0.0009 0.5 0.03
Cr 100 100 0.00018 0.23 0.075
Cu 55 20 0.01 14 2.4
Fe 56,300 38,000 0.67 140 160
Hg 0.08 0.03-0.8 0.00008 0.Q15 0.046
Mn 950 850 0.012 630 0.2
Mo 1.5 2.0 0.00035 0.9 < 0.2
Ni 75 40 0.01 3.0 0.8
Pb 12.5 10 0.005 2.7 2.0
Sn 2.0 10 0.00004 < 0.3 < 0.15
V 135 100 0.001 1.6 0.15
Zn 70 50 0.01 100 160

A traditional way of determining toxicity of metals and other toxins is to establish the LCSO or LDSO value, where
LC stands for lethal concentration and LD for lethal dose. The value is obtained for the concentration that
exterminates 50% of the test sample relative to a control group of test organisms during a certain time span. More
than 200 monographs on various toxicological test systems have been published (see, e.g., Cairns, 1981; Burton,
1994). A crude role of thumb states that the least hazardous elements appear with the bighest concentrations in
water, sediments and biota, and vice versa, the "abundance principle" (see Hakanson, 1980b). ElemenlS
appearing on the ppb (parts per billion. 109) scaie, i.e., with extremely low natural concentrations are, e.g., Hg,
Ag and Cd (table 1.2). Elements on tile ppm-scale (106 ) are, e.g., As, Co, Cr, Cu, Mo, Ni, Pb, Sn, V and Zn.
Elements on the mg-scale (10 3) are, e.g .. AI, Ca, Fe, K, Mn and Na.

The distribution form of the metal in aquatic environments is very imPOrtant for the toxicity and the potential
ecosystem effects (Gottofrey, 1990; Wicklund, 1990). Generally, the toxicity is highest for the ionic species and
proportional to the oxidation number, e.g., Cr04- is more toxic than Cr3+ The potentialloxic effects of metals
can often be significantly reduced because the metals are bound to different compounds, which may
camouflage the toxic properties. Sulphides, bydroxides, carbonates and organic complexes are examples of sucb
carrier particles.

The toxicity also often depends of tile chemical conditions of the water and tlle sediments. Low pH and high
redox potential (Eb) often increase metal toxicity. The solubility of most heavy metals increases at decreasing
pH and metals which previously have been bound in ratiler harmless particulate forms in the sediments may be
recirculated to the lake water and express their toxic properties if sediment pH or Ell changes.

7
a) ESSENTIAL (e.g. Cu, Zn)
DEFICIENT I OPTIMAL I TOXIC I l.ETHA.l.
I I I
I I I
I
I
'"
l-
~
I
I
I
0 I
a:
Cl I
I
0 I
..J I
UJ I
;:

b) NON - ESSENTiAL ( Cd,Pb)


TOI..ERA.Bl.E I TOXIC I l.ETHAl.
I I
I I
'"~
l-
I
I
I
0 I
a: I
"!. I
I
0
..J I
UJ I
;: I

METAL CONCENTRATION

ElieCI

80 .
\~§:: ~
c:(
ro a: "
~ w
"
~
w
I-
0
.0
Cl
«
60 or « >- z
0
c ;: .",

.~ ;;:
w a:
"
0- 0 ,:: 0
p
(f)
40 Z « « ..J
« w
(f)
LL
a: «
a:
w ..J :0 w
0 <{ (f) Z
0
20 Z
l-
(f)
I
(f)
~
w <(

),
w 0

,\
a. 0 a: z
0 0 LL N
,
0.01 0.1 10 100 1000 10000

(Zn), ~gn

Fig. lA. Upper diagrams. Scbematic illustration of the growth of organisms as a function of tile supply of
essential and non-essential metals (from Forsmer and Wittmann, 1979),
Lower diagram: Natural concentrations of zinc in relation to optimal conditions for aquatic organisms. From
lCME (1995).

The roles of heavy metals are complicated by the fact that some of them are essential in small amounts for the
organisms. This means that an increase in the metal concentration may simulate growth while an increase beyond
a certain limit may have inhibitory effects. A schematical illustration of essential and non-essential metals is given
in fig. 1.4A. Fig. lAB illustrates classical principles for natural limits for zinc concentrations in water in relation
to water organisms.

The major threats from chemical pollutants today concern nutrients (nitrogen in marine areas, phosphorus in
lakes and heth nutrients in brackish waters) causing different types of eutrophication effects (Ambio, 1990:
HiLkanson and Wallin, 1991; Wallin et al., 1992), toxic substances like metals (Cd, Pb, Cu, a,o.: Forstner and
MOller, 1974; Vernel, 1991) and chlorinated organics (e.g., PCBs, DOTs, dioxins; Sodergren et al" 1988) and
acidification of land and water, its ecological damage and economical consequences (Ambio, 1976; Likens et al.,
1979; Overrein et al., 1980: Merilehto et al" 1988). There are many modelling approaches to acidification of
aquatic and terrestrial environments (Eliassen and Saltbones, 1983; Sverdrup, 1985; Warfvinge, 1988). Most

8
models for acidification, however, are oriented towards chemical and physical processes for individual sites or
restricted areas of investigation, rather than for entire ecosystems (like an entire lake). There are, to the best of tile
author's knowledge, no ELS-models available for acidifying compounds applicable for defined ecosystems were
operational effect variables (related to, e.g., reproduction of key functional organisms) are quantitatively related
to time- and area-compatible load (sulphur or nitrogen fluxes or concentrations) and sensitivity variables (like
water retention time, mean depth, humic content, etc.). Many concepts that can be defined from ELS-models, like
crilicalload, were, however, originally developed in acidification studies (see Henriksen and Brodin, 1992).

Elevated concentrations of contaminants that cause no visible or measurable ecological effects would generally be
of less interest for practical water management, and to remedial strategies, in Ibe situation faced today in
ecosystems with multiple threats. The aim of ELS-models (Hi\kanson, 1991a) is to provide a tool for quantitative
predictions relating operationally defined ecological effects to compatible load and sensitivity variables.

1.1.2. The PER-concept

A critical area for environmental protection is Ibe development and implementation of general criteria for the
dimensioning of environmental problems. Formal tools are needed to establish what everybody intuitively
understands, Ibat all environmental problems are not Ibe same size. The size of a problem depends on many
different considerations, e.g., on what government level the environmental problem is discussed (global, regional
or local levels), on what time scale the problem is considered (years, Ibe period between general elections,
centuries, etc.) or what personal perspective is adopted (geographical, biological, political, environmental, etc.).
From Ibe perspective of the ecosystem scale, relatively little research has been devoted to tile important but
complex problem of developing scientific criteria to distinguish between large and small, real and imaginary
environmental problems related to chemical tirreats.

An important step in dimensioning environmental problems is to identify the effects associated with various
perturbations. If we select Ibe example of different heavy metals in fresh waters, one might assume tllat Ibe
mercury problem should have high, priority in many countries, but one should ask what criteria allow us to make
Ibis ranking? Can the Hg problem in fresh waters be compared with olber national, regional and local
environmental problems? Probably, we will never know enough to have scientifically unassailable criteria for
dimensioning all environmental disturbances, but we may be able to group disturbances into classes of different
priority. It would be of great value if we could, at least, establish a number of priority classes and accurately
explain the criteria of classification. This is tile aim of tile PER-approach.

H3.kanson (1984a, 1990) gives a more detailed Ibeoretical background to tile PER-approach (sec fig. 1.5). Many
factors may have an influence on how an effect variable varies among lakes. The effect-load-sensitivity analysis
(ELS) aims at identifying the most important factors in tilis respecl. One of Ibe advantages of the ELS-analysis is
that it provides a possibility to rank factors exerting influence on an effect variable so that future research
resources can be concentrated on these factors. Naturally, when using an analysis on the ecosystem level (i.e., for
entire lakes, coastal areas, etc.), it is not possible to explain many phenomena that occur on the individual,
organ or cell levels.

In tile subsequent parts of Ibis section, we will follow a specific scheme of addressing questions, which is
ilIustnlted schematically in fig. 1.5. The rational for tilis approach will be further elaborated with several examples
in the following pages. For eaeb problem (= chemical tilfeat, I.e., acidification, eutrophication Or contamination),
questions are first asked about:

9
PER = Potential Ecological Risk factor or Environmental Index
Define:
1. Effect variable (El. preferably linked to reproduction/survival of key functional organisms
2. Areal extent (A) of contamination
3. Duration (TI of the contamination

A. Effect-load-sensitivity model B.
(E)
E3 )
...,0 Sensitivity (S)
~
....
I'l

(L)
Load

(E) c.
remedial measure

phase of phase of
contaminatio recovery

Ei Ai Tl PERi
Emax-El Al Tl (Emax-El)*Al*El
El-E2 A2 1'2 (El-E2)*A2*E2
E2-E3 A3 T3 (E2-E3)*A3*E3 1960 1990 2200
1 00 1930
E3-Emln A4 T4 (E3-Emin)*A4*E4 .
PER = l:PERi
Duration m of
contamination

Fig. 1.5. Central concepts in the PER-approach to quantitatively rank various chemical threats to ecosystems.
A. The PER-formula shows that PER is a function of the ecological effects variable (E) and its areal (A) and
temporal distribution (T). The figure also gives an illustration of in an ideal situation where contour lines of the
effect variable are available.
B. Illustration of an ELS-diagrarn. i.e., an Effect-Load-Sensitivity diagram based on an ELS-model.
C. Illustration of the duration of a contamination in a given ecosystem with a phase of contamination,
remediation and a phase of recovery. This gives (tile integral) the T-factor in tile PER-analysis.

1. The effect variable (E).


2. TIle geographical or areal extent (A) of the effect variable.
3. The temporal extent, or duration. (11 of the effect variable.

1. The effect variable. In this system. the E-value should vary from 1 (= no effect) to 10 (; total colJapse of
original ecosystem structure. i.e., extinction of key functional organisms, like important species of fish). E = 10
means that a given chemical threat has caused a total change (100% reduction) in abundance of a defined key
functional organisms in at least one ecosystem (e.g .. a lake). The grading from 1 to 10 is problematic - what

10
criteria should be used? The choices given here are examples and they could and should be challenged! The PER-
systems can be seen as an analogy lO the well-known and practically useful Richter-scale for quantifying seismic
events.

Effect variable, the E~value:


1 = no known or likely ecosystems effects 6 = clear real ecosystems effects
2 = unlikely ecosystems effects using statistical (= probabilistic) methods 7 =substantial real ecosystems effects
3 = likely bmlow ecosystems effects using statistical methods 8 = large real ecosystems effects
4 =probable ecosystems effects using statistical methods 9 = very large real ecosystems effects
5 = small real (= martifested) ecosystems effects 10 = lOtal collapse of real ecosystems

Note that an ecosystem is an entire lake or a defined coastal area. The boundaries of a lake constitute the natural
limitations for the lake as an ecosystem in this context. However, a very large lake (larger than about 300 km2) may
have to be divided into parts where differe11l key functional groups dominate. For example, a large shallow bay may
have to be separated from the open water area. It is often more difficuitto defme the limitation for other types of
ecosystems with open andlor diffuse boundaries, like forests, rivers and coastal areas. An example how this can be
done for coastal areas will be given later.

When E is < 10, it is important to seek a "critical" andlor operational guideline value for E (Ecril, which can be,
e.g., a guideline concentration of a toxin in fish) for the determinations of the areal and temporal extent. The
motivation for such Ecrit-values are very important since it has to do with the practical applicability of the
method. Several examples of this will be given in the following parts.

2. The geographical or areal extent of the effect variable. The following classes will be applied:

Areal distribution, the A-value:


1 = no ecosystems with E = 10 or E =Ecrit . 6 = ecosystems in many regions with E = 10lEcrit
2 = a few ecosystems (here lakes or coastal areas in 7 = more than 25% of ecosystems with E = 10/Ecrit
Sweden) with E = 10 or E = Ecrit (n<25)
3 =several ecosystems with E =10/Ecrit (25-100) 8 =more than 50% of ecosystems with E =10/Ecrit
4 = many ecosystems with E = 10/Ecrit (l00-4oo) 9 = more than 75'10 of ecosystems with E = 1O/Ecrit
5 =most ecosystems in a re~ion with E = 10/Ecrit 10 = all ecosystems (in Sweden) witi, E =1O/Ecrit
Note that tile selected region for environmental management in tilis book is the country of Sweden, but this
particular selection has very littie to do witil the basic principles and processes of the overall PER-analysis or with
the fundamental components of ELS-modelling and that ecosystems here are lakes and coastal areas. TIlese
principles are meantlO be generally applicable for any country, defined region or ecosystem type.

3. The temporal extenl., or duration, of tile problem. The following classes will be used in this context:

Duration in time, the T ~value:


1 =no effects 6 = more than 20 years
2 = effects (E = 1O/Ecrit) for less tilan I montil 7 = more tilan 40 years
3 = more than 1 month 8 =more than 80 years
4 = more than 1 year 9 = more than 160 years
5 =more than 10 years 10 =effects (E = 10/Ecritl for more than 320 years

11
[n this approach, where chemicals are assessed at the ecosystem level, the toxicity is not !reated by means of
specific ecotoxicological or physiological tests (see, e.g., Boudou and Ribeyre, 1989a, b; Gottofrey, 1990;
Wicklund, 1990; Burton, 1992; [CME, 1995). The PER-value depends on defined and measured ecological effects
for given ecosystems, e.g., E1 in lake X, E2 in lake Y. PER increases when the E-value increases reflecting the
amount of virulence of a given contaminant load, but PER also depends on both the areal extent and the duration
in time. The greater the areal distribution, the higher the PER-value (then a given E1 appears in many lakes in a
region) and the longer the effect lasts, the higher the PER-value (then the given E1 lasts for N years in lake X).

How is it possible by using simple operational methods and without becoming lost in extreme complexity, to
distinguish, for example, between tolnl dose (Dtot) and biologically active dose, often approximated as the
dissolved dose (Dbio!) for a given pollutant? Which factors influence the distribution coefficient (DC =
Dbiol/DLOt)? Which sensitivity factors influence the DC-value? Which of these sensitivity faclOrs can be
influenced by remediation? How can the most probable effects on the ecosystem level be measured and predicted?
Although we may never be able to answer tllese questions fully for many environmenlrd threats, it would be
interesting to set defined goals and then see how far one can go, and in the process, to discover our limits. This
is what the following PER-analysis is all aboul

1.2. Ecosystem analyses - basic concepts

This section gives a brief background to a method to define the fundamenlrd unil, the ecosystem, as well as shon
discussions on the major environmenlrd Ulfeats nOt just 10 aquatic ecosystems but 10 life on this planel, and some
basic characteristics of management models and ELS-models. This chapter may be considered as an extended
in!roduction to the following parts of this section dealing with PER and the criteria to rank chemical threats.

1.2.1. Defining ecosystem boundaries

[n this approach, it is imponant to define the fundamenlrd unil, the lake or coaslnl ecosystem. This is not a
diffuse food-web structure, but a defined geographical area. Large lakes may have 10 be separated imo sub-
ecosystems where different key organisms dominate. Smaller lakes (area < 300 km 2) can often be regarded as
single ecosystems. If a given E-variable varies more within a defmed ecosystem (like a lake) than among
ecosystems, it becomes very difficult to develop models that predict the effect variable. Such cases may arise
where ule definition of the boundaries of the ecosystem are ecologically inappropriate. An ecosystem is a rather
uniform entity with respect to its defining characleristics. One very important question concerns tile definition of
coaslrd ecosystems in large lakes or marine areas like the Baltic. The question is where to place the boundaries
toward the sea andlor adjacem coaslrd areas. It is crucial to use a technique that provides an ecologically
meaningful and practically useful definition of tile coaslnl ecosystem.

How should one define this area so timt, e.g., map parameters, like mean depth, can be relevant as model variables
(x) to predict tile target E-variable? The problem is illustrated in fig. L6B using data on Secchi deptil (the variable
10 be predicted in tltis example; Secchi deptll is a simple, useful operational measure of water quality) and mean
depth (Dm, tile x-parameter which reflects, e.g., resuspension processes). For small lakes, tilere exists a significant
(r2 =0.38, p =0.0001 for 88 Swedish lakes) positive relationship between Dm and Secchi depth: TIle deeper tile
lake, the larger the botlOm areas beneaUl tile wave base, the less resuspension, tile less suspended materials in tile
lake water, the clearer tile water and tile larger the Secchi depth. This is logical. The mean deptil has a significant
meaning for an important ecosystem variable, Secchi depth. The entire lake is tile detined ecosystem.

12
Defining ecosystem boundaries
lAKE

Outflow

Area smeller
then = 100 km 2
::::.
Inflow
. - ·:,:.:,:.,,'.:.:.:. :.:. c...:,.:. :,.:. :.:.;".:,:·?/y·'·
Mean Secchi Exposure
depth (m) depth (m)
A Mean depth == 2 m
Area 4.5 2 0.05 Secchi depth .. 2 m
Area. B 3.5 2 0.1
AreaC 2.5 2 0.2
y -"{).25x + 5.4; r2 -0.14;n=23;p-0.08 y - 0.25x + 1.55;r2 _ 0.38:n=88;p-0.OOOI
10 10
8 8
g 6

lit. • 6 0 0
:cu ••
'0
4 • 4
u
v
rtl 2
•• ."
•• • • • 2 q,
o 0
00
0

0
0 2 4 6 8 10 12 14 2 4 6 8 10 12 14
Dm(m) Dm(m)

Minimirethee:q:osure. Ex = 100*At/Ab.
where
At - the section area (m 2) and
Ab = the 3-dlmenslonill bottom area (m 2)

Fig. 1.6. lllusiiation and rationale for the definition of ecosystem boundalies. The coastal ecosystem is defined
by the borderline marked A. which gives a minimum exposure (Ex; the "topographical bottle-neck method").

But how would this apply for a coastal area in a large lake or in the Baltic? Is tilere a method to define the
boundaries and establish coastal ecosystems where morphometric parameters like the mean depth bas a meaning
in predictive models? This is illuSliated in fig. 1.6A. In this example, there are three boundary lines. A, B and C.
The mean depths of the enclosed areas are 4.5, 3.5 and 2.5 m, respectively. Tbe exposure of tile coastal area is the
ratio between the section area and the three-dimensional water surface (Ex = 100' Ali Ab. where At =the section
area or the opening area towards the sea in lan2 and Ab = the 3-dimensional bottom area in lan 2). The Ex-values
are quite different in tile three cases, 0.05 for A, 0.1 for Band 0.2 for C, but the Secchi depth is the same in all
three cases (2 m). Arbitrary border lines can be drawn in many ways and the mean depths of such areas would be
devoid of meaning in relation to a target effect variable. The approach in this work ([rom Hakanson et al., 1986
and Pilesj6 et a1., 1991) assumes that the borderlines are drawn at the topographical bottle-necks so that the
exposure of the coast from winds and waves from the open sea is minimized. It is easy to use the Ex-value as a
tool to test different alternative borderlines and define the coastal ecosystem wbere the Ex-value is minimal. If tile
coastal ecosystem is defined in tilis way. then there exists. as illusiiated in fig. 1.6A, a weak but statistically
significant (r2 = 0.14, P = 0.08 for 23 Baltic coastal areas where the tidal impact is negligible) negative

13
relationship between Secchi depth and mean depth: The larger Dm, the more resuspension of fine sediments, the
more suspended materials in the coastal water, the more turbid the water and the smaller tile Secchi depth. This is
also logical because coastal areas are by definition open to the outside sea (the section area, At> 0; if At = 0, tilen
this is not a coastal area but a lake near the sea). For such areas, a significant portion of the materials suspended in
the water can "escape" from the coastal area to the open water area or to surrounding coastal areas. This is not the
case in the same way for lakes. So, coastal areas with small mean depths will generally bave coarse bottom
sediments with small amounts of fine materials of organic and inorganic origin causing high water turbidity when
resuspended.

It is always important to define the presuppositions of any model. When and where will it apply? The definition of
the ecosystem boundaries is one crucial aspect of this for ELS-models for coastal areas.

1.2.2. Ecosystem indices

In environmental management it is important not to use personal viewpoints as criteria to rank threats as a basis
for action, but to have a more objective approach. There is a growing awareness that much better individual
"indicators" and aggregated "indices" of environmental health are necessary because they alone could provide a
rational structure for decision-making in the environmental sciences (Bromberg, 1990; OECD, 1991). An index
(an aggregated measure) is generally distinguished from an indicator (a single variable), and an ecosystem (a
single instance, like a lake or a field) from an ecosystem type (the summation of several to many ecosystems).

This book discusses two different types of environmental indices: PER, in section I, and LEI (the Lake Ecosystem
Index), in section 4. PER is a general, holistic index Calculated from the geographical extent and duration of a
defined effects variable (E or Ecrit). LEI is calculated from changes in biomasses of key functional organisms
related to chentical remedies, which can alter these biomasses relative to defined natural (= reference) conditions.
These two systems will, hopefully, be a step to~vard a more objective platform for dimensioning environmental
problems. Certainly, the complexities involved in establishing simple, practical and meaningful ecological indices
or effect variables sometimes seem insurmountable. Still, the benefits of even crude indices like PER are so great
that they are well worth pursuing. So long as one can clearly state ones criteria, theories and evidence in these
complicated matters, then these components can be discussed, tested and improved.

A frame of reference is required to assess the status of the environment. Since 1987, many countries have
accepted "sustainable development" as a goal for environmental and econontic policy. The term was introduced in
the final report of the Commission for Environment and Development (the Brundtland Commission). However,
this phrase is empty unless it is defined in terms of operationally measurable properties, desired goals and relevant
data. There are alternatives to choosing ecosystem as the basis for environmental typology (Mackay and Paterson,
1982; O'Neill et al .. 1982; Cairns and Pratt, 1987). Instead, one might use different geographic areas or different
media such as air, water and soil. There is, however, a clear international trend towards consideration of tile
"healtil" of the different ecosystems (Bailey et al., 1985).

14
1.2.3. Environmental threats

Tbe environmental threats to life on this planet include:

Chemicals involved
I. Climate change C02, etc.
2. Reduction of the ozone layer 02, freons, etc.
3. Acidification S,N
4. Air pollution S, NOx, Pb, etc.
5. Eutrophication P, N, etc.
6. Contamination of metals & radionuclides Metals and radionuclides
7. Contamination of organic toxins DDTs, PCDs, dioxins, etc.
8. Health effects and inconveniences CO, Pb, S, etc.
9. Changes to the rural landscape areas worthy of protection Xenobiotics
10. Reduced biological diversity Xenobiotics
II. Introduction of exotic and new organisms
12. Over-exploitation of natural resources

Ten of these twelve threats involve chemicals. A set of ecological effect variables is expected to reflect such
threats and the extent to which they affect the ecosystem. Note the difference between biological effects for
individual animals or organs and ecological effects for entire ecosystems. Practically useful, operational effect
variables should be:

measurable, preferably simply and inexpensively


clearly interpretable and predictable by validated quantitative models
internationally applicable
relevant for the given environmental threat
representative for the given ecosystem.

The effect variables or indices must be chosen so that the "distance" between the present environmental status and
an identified environmental goal can be determined. Ideally, environmental effect variables should be
comprehensible without expert knowledge. In fact, one reason to develop such measures is so that politicians and
the general public can understand the present condition and future changes in the environmenL

The creation of an ecosystem index like PER requires aggregation of information. For example, if the indices for
all ecosystems in a region are averaged, this figure is then a regional ecosystem index. A still higher level of
aggregation is obtained if one sums (or averages) the regional indices for each ecosystem type (for lakes, forests,
agriculruralland, etc.) into a single regional or national environmental state index. An aggregated index of
environmental health would complement the picture of the country's economic development given by the GNP
(gross national product). An environmental state index of this kind could be compared with a consumer price
index; environmental indicators would correspond to different items, the national environmental index for a given
ecosystem type would be similar to the value of a class of goods.

Ecosystem indices would have the advantage of expressing the environmental status simply, but they
simultaneously pose problems in that a great deal of valuable information is lost in aggregating the individual
measures. This disadvantage is reduced if one knows exactly what an index represents, and if one can access these

15
individual components as required. Ideally, Ibe same basic framework would be used at bOlh Ule national and
regional scales. However, since problems and priorities cannot be completely congruent at different levels, Ibe
framework may be adapted to Ibe different requirements of different levels. Tbe national level may address large-
scale Ibreats, perbaps originating outside Ibe country, like acidification of soil and water, whereas Ibe region can
address more local problems, like Ibe eutrophication of lakes.

A crucial question is: How could Ibis be achieved in practice? The following parts of Ibis book will give one
avenue to this very difficult goal.

1.2.4. Target ecosystems

An environmental index must be based on Ibe status of some crucial characteristics of chosen ecosystem types,
sucb as:

1. Forests
2, Agricultural land
3. Natural land
4. Freshwater
5. Coastal areas and
6. Urban areas.

These are the six basic ecosystem types.

As pointed out, it is extremely difficult to distinguish cause and effect in natural ecosystems. One cannot base Ibe
PER-number or the ELS-model on a full understanding of Ule ecosystem. In complex ecosystems "understanding"
at one scale (e.g., Ule ecosystem scale) is generally related to processes and mechanisms at Ibe next lower scale
(e.g., the scale of individual animals andIor plants), and the-explanation of phenomena at this scale is related lO
processes and mecbartisms at the next lower scale (e.g., the scale of the organ), and so on down to the level of the
alOm and beyond. In environmental management, the predictive ecologist must often find a balance between
answering interesting, often important, questions of understanding, and delivering a practical tool to society. If an
ecosystem index were based on a causal analysis of what takes place at the cellular level, then at levels involving
organs, individuals, populations, and finally at the ecosystem level, one would wait an eternity before the index
could be developed. For the foreseeable future, ecosystem indices like PER are more likely lO be based an
practical considerations of predictive power and sampling ease, rather than full causal priority.

What, then, does the strategy look like if one wants to develop an ecosystem index like PER in practice" The ftrst
problem is that each ecosystem type, e.g., fresh waters, is not a single entity. It consists of many sub-ecosystems
(fig. 1.7). A general resolution about the base of this approach is probably impossible, but questions about Ule
appropriate hierarcbicallevel of analysis are relevant to specific Ulfeats. Fig. 1.7 lists Ibe 12 general
environmental Ulfeats mentioned earlier. If one starts with the threat to fresh waters, it is clear that contamination
of, e.g., metals/radionuclides Ibreaten fresb waters and that these Ibreats might be manifested in, e.g., reduced
biological diversity and contaminated fish. It is also clear that some of these 12111rcats are not relevant to fresh
waters. "EveryUling does NOT Ibreat everything elsc"!

16
Fresh water Ecosystem

Soil water Surface water

In-flowareas l0ut-flowareas\ Lakes

Target ecosystem
for metal
Mesotrophic Eutrophic
contamination

t tt t t t t t t t t
.g"
u ~
'0
c:
~ '"
'" '"0.
0 ~ ~

"» '0 s
0
~ "
'(3
~ '" u '0 1l. ~

"c: "
-5 .~
.!l
u b
~
::"
:a""
S .!l
0;
'" 0 "§ "0
0
u "
c
<!'
0
~
0. ,." "
~

.=
~
'1: :a '0

.a
0
N .S! OJ
"0.
.c 0 >< ~

"
~

. .9 2
'"
0
~
0 ~ "0. u
'8,
"t: "t:
s '"
,. " .0 '"
U
~
0
"t: 01 OJ t:
0 'c
"
0 !<l ~
~
0
~
0
~
.!< "g " 'c u ] J3 t:
"
"u
.c 0 .S!
.c '" <ll 0 0

-
'0 u '0 'c
.S"" "'" S g. t:
.~
~ '0
"~ "B E
.!l " "~
g" "0
.c
'"u ]
.5" "
~
c I:
-< "'"
!:tl "
0
.;;;
:r:"
.!l
:;;:
l'
u
0"
'0

.s
C.
><
0 ~
" 4
u !:tl

t
1 2 3 5 6 7 8 9 10 11 12
12 environmental threats
ELS-models available
for Cs-137 and Hg

Fig. 1.7. The fresh water ecosystem may be divided into several sub-ecosystems (such as oligotrophic,
mesotrophic and eutrophic lakes) where different key organisms prevail. As pan of this differentiation process,
one must decide which sUb-type should be used as the target ecosystem for a given chemical threat, which
target organisms, functional groups or effect variables should be used for a given threat, and what load and
sensitivity variables should be used relative to a given effect variable. The target ecosystem for metals (like
Hg) and radionuclides (like Cs-137) is low-productive (oligotrophic) lakes (shaded in this figure).

17
A Test results for different biological organisms
Example: Cadmium
- T arget organlsms
Long..term effects .-~\J
.- ! effects
1// Short-term I
---
Concentration In water: I4l Cdfl

B. Results for dJfferent fresh water ecosystems

Example: Cadmium Target ecosystems


Long-term effects I", L //
on target organisms" V'e 0
;a
In the given ecosystem
1 ~ a.
]o ..-
0
b
~ o , 0"
"
~ -a e "'-a
0_
::l_ '"
1 2 4 8 16 32 64 128 256
Ecosystem Index
Fig. 1.8. A. Illustration of target organiSms. Data on effects of cadmium in sbort- and long-term laboratory
tests (SNV. 1980; see fig. 1.9). The most sensitive organisms in long-term tests are the most appropriate target
organisms. Tbe tests may indicate the weakest link in the eco-cbain and the concentration at wbich that link
breaks. but they do not sbow that this organism is, or is not., critical to the ecosystem [unction.
B. An illustration of the identification of (hypothetical) target ecosystems according to the sensitivity to Cd-
contamination.

One way of relating the threat to the "weakest link in the eco-chain" is illustrated in fig. 1.8. Tbe metal cadmium
is used as an example, but the principles of the discussion apply to other contaminants as well. Various studies
(mainly in the laboratory) show that certain species are sensitive to cadmium and that others are not. In the case of
the most sensitive species, cadmium concentrations in the water o[ 0.1-10 fig/I disturb reproduction. If these
sensitive species are key members of the ecosystem, such as top predatory fish or feed organisms for such fish.
then cadmium concentrations as low as 0.1-10 figlI will damage the ecosystem. The sensitive species will be
decimated and tile original balance among naturally OCCurring species would be changed. Furthermore, the
concentrations of Cd, and most non-lipophilic forms of metals. are often, contrary to common belief, lower in
biota than in abiotic environments, such as sediments. Generally, for Cd and other metals, concentrations appear
in the [ollowing sequence: water (lowest), fish muscle. fish liver. algae, macrophytes, insects, bivalves and
sediments (highest; see fig. 1.9; see also fig. 1.21). The high sediment concentrations mean facilitated laboratory
analysis and more reliable measurements, as compared to water samples and many biological samples. Tills is an
important aspect in water management and monitoring. A target organisms for an ELS-model for Cd would then
be selected from the most sensitive species.

18
~
%
9
0
%
o •
!~
w
0%
%0
0-
!
0';
%w
O~
•% •.
%
0
~
~
, !< 5~
=,
~%
<0 %
~% 0
-<
!;(?;
9
~
~
<
W
~.
~
w.
~
~ ~ ~ w
~<
~m
~ - ~
~
~
~

u< E uU
w %. %%w %%
% -0
~u
,~ ~~
'u ffi !< w
U W
Ww w~
u -~.
u
w.
w o . u-~
~w Ow
%~ \i~ % • %0
w %w . % % ~ W

*~
0 o·u
0 o. ou %_::.
o~
.w
0
u ! u ! u< >
W
u ~n;; 8 :: i

!
~ 10'
i I
e

c
E
·•
! 10·

·
;
,

Q
I
Q

"-
~ 3
;; 10
• ·
~, ·
102

··
Q
~
I!
ii
0
~
·
. I
~
~
%
w 10'
U
I'
%
0
u ·
·
I
I
I
••
I'
I

10 0
P90
P15
I I
Pso
P25 M

,o~

N:~32
\1 I
I

10~
i I
Fig. 1.9. Measured Cd-concentrations yielding shon- and long-lenn lOxic effeclS in laboratory leslS. and Cd-
concentrations in river water, lake sedimenlS, a1l2ched algae. macrophytes, invertebrales, fish muscle and fish
liver (dry weighl of organisms pUl at 20% of wel weighl). Based on data from SNV (1980), figure from
Hilkanson and Jansson (1983).

19
pH (dose) vs Ecological/biological effects
7.5 5.5 5.0 4.0 3.5

White moss
and filamentous algae
increase

7.0 6.5

CrustaC'""...ans. molluscs, etc. dle

~ I Target indicators I

5.0 4.5 4.0 35

Salmon. char. trout. whitcftsh


"",vlino. perch and pike die

Fig. 1.10. Illustration of target organisms using the example of ecological effects of lake acidification. In
predictive limnology. ecology and environmental sciences. a central objective is to frnd out whicb organisms
are the most sensitive to a given contaminan~ whether this is an acidifying substance (like S or N). a nutrient
(like P or N) or a toxin (like many metals. organics or radioisotopes). Tbe figure sbows examples of target
organisms for acidification. Crustaceans react rapidly to changes in pH. whereas cenain fish sucb as brook
trout and eels do not die until acidification is far advanced. Wllite moss (e.g .• Spagnum) and filamentous algae
sbould not be found in these lakes under normal conditions. So. the abundance of such species also indicate
ecological effects of acidification. This figure raises many questions: Since pH is a variable. wbich measure of
pH sbould be used for which organisms? Since all organisms vary spatially and seasonally in a lake, bow could
indicators of the effect be operationally defined? What would they represent? From Hiikanson and Peters
(1995).

Another way of illustrating the important concept of target effect variables can be taken from lake acidification
studies. The pH of the lake water is important in many contexts (fig. 1.10). For example. many animals
accustomed to a circum-neutral pH (= 7) cannot reproduce and survive in acidified lakes. Some. like crustaceans.
snails and molluscs. are very sensitive to cbanges in pH, whereas other animals, like salmon and pike. are nol. For
crustaceans. the most important pH-value may be the lowest pH during the spring-flood wben Uley reproduce, and
for the mercury content in pike, it may rather be the mean. long-term (2-4 year) lake average pH. TIlese typical

20
examples apply to pH, but the same questions about representativity and time·compatibility could be raised for
any x·variable expressing threat in relation to any ecological effect variable.

The target organisms are the most sensitive species in the ecosystem, but their roles may be critical in only some
ecosystems, e.g., in oligotrophic lakes. Such lakes are then the target ecosystems for this threat. The target
organisms might not be present or might not have the same key function in more productive systems, so such
productive systems may "withstand" higher contaminant loads than low.productive (= oligotrophic) systems. This
illustrates the term environmental sensitivity (= vulnerability) - one and the same contaminant load may cause
different environmental effects in ecosystems with different sensitivities. Consequently, it is important to define
both target organisms and target ecosystems relative to a given chemical threat. This prinCiple applies to all
environmental disturbances, but frequently it is very difficult to apply in practice. When this is the case, it is
particularly important to distinguish what we should know from what we know and wbat we do not know. To
clarify such matters is an essential part of the PER·analysis.

1.2.5. Regression models and mass-balance models

Differential equations are often used to handle fluxes (e.g .• g X /yr). amounts (g X) and concentratious (g
Xlm3 ) of all types of materials (like gases. carbon. toxins and nutrients). but not ecosystem effect variables (E).
Regressions based on empirical data are generally necessary to relate concentrations of chemicals to target effect
variables (E). In theory, both these model approaches (see fig. 1.11) may be used for the ELS·analysis. provided
that at least one operationally defined ecological effect variable relevant for the load variables(s) in question is
included in the model. Ideally. the E·variable should express the reproduction. abundance, mass or status of a
defined key functional organisms (preferably fish at the top trophic level), which characterise the given
ecosystem. Such ideal effect variables can, as already stressed, NOT normally be predicted with differential
equations and mass·balance models. One classical way to circumvent this problem is to use dynamic mass·
balance models to handle concentrations of pollutants and empirical models (like regressions) to link tilese
concentrations to effect variables. If such ideal ecological effect variables cannot be operationally defined for
practical, economic or scientific reasons. then one should tr} to seek simpler but relevant alternatives, like mean
concentrations of given toxic substance (·ces) in predatory fish.

Fig. 1.12 gives the principle components of a general ELS·model illustrated as a ELS·diagram. Environmental
goals (generally set by National Environmental Protection Agencies) should be related to ecological effect
variables and not to load variables, since one and the same load may cause very different effects in ecosystems of
different sensitivities. The figure also gives some of the ELS·variables tha~ in the future, may be used for metals
in lakes or coastal areas. From this diagram, importanl concepts like natural background concentration, critical
load and environmental goal (linked to defined effect variables) can be scientifically defined. There are, to tile
best of the author's knowledge, no practically applicable ELS·models for metals today (except for mercury and
radiocesium). This means that there exist ample room for speculations about cause and effect, and about the best
strategies to remediate aquatic systems polluted with metals.

The primary interest is not on site·specific conditions ("the sampling bottle"), on the indiVidual, organ or cell
level, but on the ecosystem level, which is the perspective that should be of main interest from a management
point of view where questions are posed concerning tlle status of larger water bodies (ecosystems/environments),
and the remedial actions that could be used in practice to improve the conditions in such systems.

21
Mass-balance model Effect-load-sensitivity model
Amounts. fluxes and Ecological effects for entire ecosystems
concentrations

EnVironmental
sensitivity
Input valiable
load or function

gOCIn Change in sensitivity may


change the ecological effect
variable

Change In
e1n means Ecosystem load:dose Change In load. ein or C,
change In C. variable or function may change the e<:Olog1cal
effect variable

Fig. 1.11. Illustration of the fundamental difference between dynamic, mass-balance models and ecomeuic
models, i.e" effect-load-sensitivity models. The three wheels indicate mat by means of remedial measures one
may reduce the load variable in dynamic models and me load and me sensitivity variables in ELS-models.
From Hakanson and Peters (1995).

The predictive accuracy of the model should always be stated (see section 2.4). The predictive power is
generally determined by means of statistical tests, where model predictions of E are compared to reliable
empirical data of E. Models mat provide r 2-values (r2 = the coefficient of determination; r = me correlation
coefficient) higher man about 0.75 (and p-values lower than 0.05; Prairie, 1996) could generally be used in
practice in management for predictions in individual ecosystems (see section 2). Models providing lower r2_
values (but still p-values < 0.05) could be used for regional predictions, when predictive failure in individual
ecosystems could be accepted (Hiikanson and Peters, 1995).

It is generally not possible to derive ELS-models which apply wim equal success for all types of ecosystems. This
means that me operational range, the domain, of the model must be explicitly given to avoid abuse of me model
for ecosystems for which it was never intended to be used.

If dynamic (time-dependent) mass-balance models can meet mese requirements, mey would generally be
preferable to statisticalJempirical models because of me fact tllat mey can provide betler understanding of
mechanisms and processes.

22
ELS-DIAGRAM SENSfITVITY
• Coastal morphometry: size. form & special
parameters (surface & bottom water retention
FOR DEFINED ENTIRE ECOSYSTEMS time. bottom dynamic conditions).
• Nutrient stalUS (different forms of P & N).
• Humic influences (Fe, Mn & colour).
• Temperature (salinity & stratification).
• Water chemistry (conductivity, AI. etc).
• Oxygen conditions (02-conc. & redox
pctential).

EFFECT
• Me, PCB & DDT cone. E
in fish muscle & liver F
• FISh growth factor. F
age/weight E
• EROD activity & C
cytochrome P 450 T
Low
activity in fISh liver

Environmen!al
goal

Critical/oad LOAD

NallUal backgrowui value


LOAD
• Me, PCB & DDT cone. in
water & sediments

Fig. 1.12. The basic set-up of an ELS-diagram from an ELS-model. This figure also gives examples of effect.
load and sensitivity parameters that may be used to derive ELS-models for metals, PCBs andlor DDTs in
coastal ecosystems. Based on Hakanson (1994).

1.3. Acidifying substances

1.3.1. Background

The objective of this section is to give a very brief overview of lake acidification. The literature on acidification of
land and water, ilS ecological damage and economical consequences has grown exponentially. Important
publications include Octen (1968; Svante Oden is often referred to as the "father of acid rain"). Likens et al.
(1979). Ambia (1976) and Monitor (1981,1986). There are many models and modelling approaches to address
the natural and anthropcgenic acidification of aquatic and terrestrial environments and to propose remedial
measures for anthropogenic acidification and even entire books of literature references on acid precipitation (e.g .•
Seip et al .. 1989). In Sweden. lake liming (see section 4) has become a major industry; about 8000 of Sweden's
lakes have been limed.

23
Sall processes
Podzal profile degre.dll.uon
Utter humIc ma.tter .. HCO 3
'::l TransItional horizon (A)

~-------
Eluvla.l (bleached)
weathering
horl:::on (E)
sillclltes .., dey minerals
nllture.l ccidliiclltion
ntuviol (B) horl:::on prOCe33es

Plll"ent ro::k
material (C)
• Besic rocks: Cc.~rich. easily weathered (e.g .. dle.base)
lOCk!.,
pH • Acid rocks: Cc.~poor. slowly weathered (e.g .. granite)
cultlv..uci a1k ·lntermedlc.te rocks. e.g .. sedimentary gne1s3
h.nd..C'I:(. cond
HQrtr;ont .. l CaMg
~ tu.nspon
V.mea.!
tunspon
JjydroJog~procys~~S
So~
loy .. Ik '-..:5.11",/1003 in water

Fig. 1.13. Major processes and routes of transport of natural and anthropogenic compounds affecting tbe
acidity of rivers and lakes. From Hakanson and Peters (1995).

Fig. 1.13 illustrates tbe sources of acidification. The burning of different fossil fuels (coal, oil, petrOl, ete.) results
in emissions of different types of compounds (S02, NO x , etc.), whose atmospheric oxidation ultimately gives rise
to deposition of H+, i.e., to acidification. Acidifying substances, which are mainly sulphur and nitrogen, are
deposited on land and water as wet and dry deposition. The latter is oft.en very difficult to quantify. The acidity
of lakes and rivers is influenced by many natural processes, especially tbe degradation of organic matter and tbe
weathering of different minerals. Weatbering rates and tbe acidlbasic properties of the runoff water depend on the
chemical characteristics of the parent rock. Basic rocks, like limestone, are generally rich in ea and easily
weatbered. Acidic rocks, like granite, are neitber.

WheU,er both natural and anthropogenic factors influence tbe acidity of fresh waters was the subject of intense
debate in tbe late 1960's and 1970's. Mass-balance calculations based on extensive field measurements (Eliassen
and Saltbones, 1983) have provided a quantitative basis for ranking different processes. There is no longer any
serious doubt tbat antbropogenic acidification is a major cause of severe problems in many counties.

1.3,2. The geographical perspective

Figures 1.14 and 1.15 highlight several hnPOrtant facts:

• Acidification is not a local environmental problem of short duration (Granatb, 1980; Dickson, 1986). It is a
major, regional-to-global problem linked to tbe burning of fossil fuels and intimately associated \Viti, modem life
and technology.

24
Sources of uncertain Origin 140
To other countries and the seas
140

From FInland 10
;'
From Norway 10 .... To Finland 10
-·I~m
To Norway 10

...
. . . . . From BalUc area 10

'\ ~~

From UK40
I
I
i
1\,
;. \
\~
~
~
.... . ..,
I Frorn w~ '- From Poland 20 To Baltic are..'l40
Gennany \ From E.
I 30 \ Germany
l"rom rr'tnce: 10 30
\
From Czechoslavakia 10

Fig. 1.14. Acid rains cross many borders: Sweden's sulphur exchange with other countries in 1978
(kilotons/year). 300 tonnes of sulphur were emitted from Sweden, 100 remained in the country, the rest was
transported to other countries, mainly to Finland. Redrawn from Monitor (1986), from Hakanson and Peters
(1995).

8 - \o;JIha yr

• VII burning
rn Brown coal 6-

o Coal burning 1------:


4-
No,-N,

2-

- '" '"
000

~ ~ ~
...
0

~ 0- , , , ,
1955 60 6S 70

75 •
80

85

90
Year

Fig. 1.15. Left panel: Quantities of sulphur emitted from burning fossil fuels (based on data from Overrein et
al., 1980; Blagul mi1ja, 1982; SNV, 1982; Monitor, 1986).
Right panel: Wet deposition of sulphur and nitrogen in central Sweden 1955 to 1990. From Monitor (1991).

25
pH In small rivers
Acidification D >5.4
Sulphur and nJtrogen
Ix<'),:;1 5-5.4

GeographJcal
perspective

pH (dost:) VB Ecological/biological effects

~,m:llil;:,mollUbPlO.etc.dm

Fig. 1.16. A map of Sweden illustrating the lowest measured pH in small rivers (1971 to 1985) and lake pH
associated with different ecological effects (based on Monitor, 1986). From Hiikanson and Peters (1995) .

• Acid mins cross all national borders (fig. 1.14). They affett most types of ecosystems, and most branches of the
environmental sciences .
• Since 1860 emissions of S02 bave increased from less than 20 million tons per year to about 400 million tons
per year (fig. 1.16, left), and are mainly due to the burning of coal. Fig. 1.17 gives a map of the wet deposition of
sulpbur in Sweden during the 1970s. The deposition was then, and still is (see fig. 1.16, right), very bigh in the
southern part of ille country and decreases from south to north. TIle load of sulphur is high in Sweden, and ille
sensitivity to acid min is also very high since the soils are generally morainic with a low buffering capacity. The
exceptions from this general rule are ille counties Ski'me and Gotland, wbicb lie on calcareous bedrocks, the rest of
the country is dominated by acidic bedrocks like granite and gneiss. There has been significant and costly
reductions in tile emissions of sulphur (see below) from Sweden (see Monitor, 1991), but ille trend (fig. 1.16B)
still shows increasing deposition for nitrogen and only a slow decrease for sulphur because the emission
reductions for sulphur in other European countries have been much smaller than in Sweden.

26
Wet deposition
of sulphur in
Sweden at the
end ofthe 1970s
ingS/m2*yr

Fig. 1. 17. The wet deposition of sulphur in Sweden at the end of Ule 1970s. Redrawn from Ministry of
Agriculttue (1982).

Sulphur
1980 1990 Reduction in % between 1980 and 1990
Sweden 232 97 58
12 European countries 20,252 15,931 21

• The burning of fossil fuels also results in Ule emission of many types of metals, like Hg (see fig. 1.32, later on) .
• The problems associated with fossil fuel emissions (e.g., acidification of lakes, rivers and forests, eutrophication
of the sea and metal contamination) can only be stopped through international co-opemtion.

A list of the ecological problems that acidification produces in lakes and rivers is very long. Fig. 1.16 gives only
some important examples. The map in this figure illustrates geographical extent. The pH in not just a few rivers
and lakes, but in entire regions has been significanUy altered by acid min. The map of the lowest recorded pH in
about 4000 small rivers from 1971 to 1985 shows that large areas have pH values below 5. This means tilat many
key organisms are extirpated (fig. US, right). The E-faclOr is certainly 10. Of tile 83,200 lakes in Sweden,
about 16,000 have been drastically altered by antllropogenic acidification (Henriksen and Brodin, 1995).

27
Target ecosystems and target indicators

Entrophic

Lake.

Fresh water

Acidlfication

Fig. 1.18. Illustration of target ecosystems (shaded boxes) for fresh water acidification (ule environmental
threat). Ecologically critical organisms (also in .shaded box) living in these systems are potential target
indicators. From Hakanson and Peters (1995).

Acidification is, however, not a threat to all aquatic ecosystems and all organisms (fig. 1.18). Acidification
usually causes great damage to ground water systems (wells, etc.) and to oligotrophic lakes. Eutrophic lakes and
marine ecosystems can withstand acid rain much better due to their better buffering capacity. Thus, groundwater
ecosystems and oligotrophic lakes are target ecosystems for acid rain research and remedial measures.

1.3,3. Acidification and metals

One well documented consequence of soil and ground. water acidification is that many metals are dissolved at low
pH and Uillt elevated metals concentrations have been measured in waters of wells, rivers and lakes (fig. 1.20,
right). 111is is serious for toxicological reasons since metal toxicity and biouptake is gencrally much higber for
metals in ionic forms than for metals bound to carrier particles or in solid form (fig. 1.19). Especially cadmium
(see tig. 1.9) can, in acidified soils and waters, reach concentrations above the criticallirnits. Thus. the leaching of
many metals from the organic soil layer is highly pH·dependent (fig. 1.20, left).

28
>< ~Ion
Complex

~
InorganiC Dissolved Olelate
Molecule
1. Metal form Organic PartiCUlate~
Colloid
"Carrier panicles" Sorption
Suspended

_______ Synergism
2. Presence of other substances _______ Addition
Antagonism
Temperature
pH

~
GeOgrnphical position Dissolved oxygen
3. Water system Geology and drainage area ~:::::::=-_ Light
Morphometry of basin
Hydrological conditions

Trophic level
Humic level
Water retention time
Bottom dyuantic conditions
Temperature Sediment types

~~~
~sex
4. Biological system ~~====~--- Food supply
~Feedbabit
~ACtivity
System of protection
A~ptation to metals
Ecological status

Fig. 1.19. Factors affecting the toxicity of metals in aquatic systems. Modified from Hakanson and Jansson
(1983).

Note also mat very higb lake concentrations of aluminium (in fig. 1.20, right) appear at low pH. Aluminium
works in more or less me same manner in lakes as in water purification plans, where technical AIS04 is added to
increase flocculation and sedimentation of pollutants in waste water. If the AJ·concentration reaches above 100-
200 I1gl1. fish gills are clogged by aluminium flocs and the fish die from suffocation. So, AJ is not toxic in the
same sense as, e.g., Cd (see Pan, 1983; Block, 1991), but the end result of high Al-concentrations in lakes is tilat
fisb, i.e., key functional organisms, die.

The importance of identifying larget indicators for metals is illustrated in fig. 1.21. TIlis figure schematically
illustrates, on the Ie ft, that effects of the metal zinc can be obtained in a very wide range of concentrations in
short-term and long-term laboratory experiments with various types of biological indicator organisms. The most
sensiti ve indicator organisms, mose in focus here, react to Zn-concentrations in water of about 100 ppm, whereas
the most zinc-resistant indicator organisms do not react until the concentration ofZn in water is 3-4 powers of ten
higher.

29
AI
~gJ1
100

600

500
Release


400

• •

300 •

• ••
• •.
·••••
••

•••

100
•··• ....

• • •••
• ..
•• • •
..........
: •
.::.:::- • •• •·
4,0 5,0 6,0 1,0 .,0 pH
2 3 4 pH

Fig. 20. Left The leaching of metals from soil systems increase with decreasing soil water pH. From Ministry
of Agriculmre (1982).
Right: Al-concentration in lake water versus pH. The critical limit is about 100-200 I-ig Al per litre. From
Blagul miljo (1982).

The right side of fig. 1.21 illustrates that Zn-concentrations' in natural waters are generally relatively low. that the
Zn-concentrations in sediments are often comparatively high. and that zinc does not bioaccumulate or become
biomagnifieci. but in contrast Zn-concentrations (like most heavy metals) become lower when one moves from
bottom fauna to fish. However. the concentrations of Zn and many metals are often relatively high in specific
organs like fish liver since the liver and kidneys of fish function as the detoxification centres of the body.

1.3.4. The time perspective

For how long will acidification continue? Since the burning of fossil fuels is the basic cause for anthropogenic
acidification. and since people are likely to continue to use oil and coal. the acidification problem will probably
continue for many years to come. The problem will not go away until the aunospheric emissions of sulphur and
nitrogen have been drastically altered.

An interesting time-scale to the problem is given in fig. 1.22, From this figure, one can note that the acidification
in this lake. Lake Gfu'dsjon in western Sweden. staned around year 1960. so it is of late date. and the problem is
likely to prevail for centuries!

30
Short-term
10 6 effects Zinc
10 5
Long-term I
104 e:flects SedI-
ments
AI. AI. F1sh Uver

/~
103 AI.

10 2

J 101 TARGET
INDICATORS
100

10- 1

Fig. 1.21. Left. Compilation ofZn-concemrations (ppm) yielding sbort-term and long-term toxic effects on
various biological indicator organisms in different laboratory tests and illustration of Lbe target indicators.
Rigbt: Concentrations of Zn in natural waters. sediments. algae. bottom fauna and fisb (ppm. dry weigbt).
From Hakanson (1990).

7 6 5 6 5 4
pH
I , I I
I I I I

--
1~79
....
t
197~.
1196,0 ••
"
• ••
• •

• " •

• •



• 12,500 B.C•
,
• •

Fig. 1.22. pH-development since 12.550 B.C. reflected in Lbe sediments of Lake Gilrdsjon. western Sweden.
Redrawn from Ministry of Agriculture (1982).

31
1.3.5. Summary - acidification

Utilizing the criteria to rank environmental threats, as given by PER one can note that:

• Anthropogenic acidification kills key functional organisms. The ecosystem structure of 10.000 to 20,000
Swedish lakes have been totally altered by anthropogenic acidification. The E-value is certalnly 10.
• Acidification has a very wide geographical spread. The A-value for Sweden should be either 6 (E = 10 in many
regions) or 7 (25% of 83,000 lakes with E = 10). Using the precautionary principle, it is set to 7. Note that these
values only apply to Sweden. Acidification is a very severe problem in many other parts of the world, e.g.,
Canada, Norway, Finland, Russia and northern U.S.A,
• Acidification is rooted in the use of fossil fuels and will probably continue for centuries. The T-value is therefore
set to 10.
• PER is very high (= 10*7*10 = 700) for acidification, which is the worst menace of all the chemIcal threats to
aquatic ecosystems in Sweden according to this approach (see section 1.7 for a compilation and a quantitative
ranking). PER = 1000 means that all ecosystems of the given rype in the country (or the defined region) will have
a 100% reduction of key functional organisms for centuries.

1.4. Metals

1.4.1. Brief background on metals

Many metals, like Cu. Cd, Hg, As, Cr, Ni and Sn used to be regarded (Forsmer and Muller, 1974; Forsmer and
Witunann, 1981; Salomons and Forsmer, 1984; Monitor, 1987; Lithner, 1989; Lindestrom, 1991) as potentially
hazardous to aquatic life. During the last years, however, one can note that this is not generally the case (see, e.g ..
Lindestrom, 1991). But all these beliefs and assumptions that certaln metals were harmful or harmless have rarely
been documented by solid empirical investigations of real aquatic ecosystems, but rather deduced from controlled
laboratory experiments andlor site-, time-, and species-specific studies in restricted areas. It is a regrettable fact
that today there exist no validated ELS·mooels for metals (except for Hg and radiocesium) or for halogenated
toxins (like PCBs and DDTs). This means that there is room for speculations concerning the ecological effects of
these contaminants as such, and especially in real situations were many different substances contaminate at the
same time and antagonistic and synergistic effects can appear. Only empirically validated ELS-models provide the
prerequisite for quantitative simulations in aquatic management so that practically feasible remedial measures can
be discussed and the consequences of such measures predicted; thus enabling quantimtive environmental cost-
benefit calculations to be made (Hiikanson and Peters, 1995).

This section will use the PER-analysis and focus on one stable metal, mercury, and one radionuclide, radiocesium,
as type elements. Basic principles and processes regulating metal transport and ecosystem effects have been
discussed by, e.g., F6rsmer and Witunann (1981) and Monitor (1987). Mercury and radiocesium represent twO
different types of contamination, a continuos one, like that of Hg, and the sudden peak for radiocesium after thc
Chemobyl accident of April/May 1986. The difference that Hg has an organic/Jipopbilic phase, which Cs-137 has
not, and the differences in retention times in fish between the two elements (years for Hg and months for Cs); see
Hakanson (1996a). A dynamic (i.e., time-dependent) ELS-mooel for radiocesium will be discussed in section 3.
An empirical ELS-mooel for mercury is used in this section and presented in more detail in section 7. A dynamic
ELS-mooel for Hg, derived from the empirical model, is presented in section 4.

32
...'I ".
I T . I I
Nutrientsl .1 Acidifying Toxic Complex
Eutrophicatio~ substances substances waste water

I ... ,... I"


I
~
'I . I
Chlorinated
organics Metals RadiOisotope:

...,. ....,. It- y .....


Cr
...
Cu Hg Ph Cd
• • •
."..
tHect varia
hie known?

I I
....
No Ves

T
Mesocosm
tests Hgpi

I
T t
Hazardo\Js I
rUect
variable concentrations

,I.
Fig. 1.23. A chemical threat to a lake system can ftrst be divided into effect categories such as eutrophication
from nuUienlS nitrogen and phosphorus, acidiftcation from' sulphur and nitrogen, contamination from various
substances andlor from complex\vaste waters and then into further sub-groups. Hgpi is Hg-content in I-kg
pike, a standard operational effect variable for mercury contamination in lake management. From Hilknnson
(1990).

1.4.1. Mercury

The flow diagram in fig. 1.23 is meant as a "navigational tool" to facilitate the environmental consequence
analysis. It consislS of a number of boxes for different aquatic environmenlS, rivers, lakes and marine areas. If the
discharge takes place in a lake, then one proceeds to the next level in the system. Here, there are three
possibilities: Is it a question of discharge of nuUienlS, which may lead to eutrophication, acidifying substances,
toxic substances or complex wastewaters? In this section, we will treat mercury.

1.4.1.1. Effects variables for mercury in aquatic ecosystems

Ever since the Minamata disaster in Japan in the 1950s, mercury has certainly been one of the most studied
environmental pollutanlS. There are about 60,000 publications dealing with mercury as an environmental
pollutanL But contrary to common belief, Hg does not appear to be a major threat to aquatic life. This facL that it
is difficult to find environmental effeclS of mercury in aquatic ecosystems (but there are effects for, e.g., birds),
and of other heavy metals in natural waters as well, depends, for example, on the affinity of metals to become

33
bound to different types of canier particles (e.g., clay minerals, humus and algae), whereupon the biological
uptake of the metals is more difficult. Thus, mercury does not constitute any known threat to aquatic ecosystems.
However, it is certalnly a threat to the foetuses of pregnant women if the mother eats Hg-contaminated fish. The
operational effect variable, which has been used for Hg for many years, is the Hg-concentration in I-kg pike
(Esox lucius). Pike is a stationary predator eaten by humans. Evidently, this is not an effect variable in the same
way as reproductive disturbance in roach or mortality in crayfish due to acidification (fig. 1. 10). However, it is the
target operational effect variable for mercury in lakes. Today, this can be established without, e.g., mesocosm
tests (fig. 1.23).

The long-term environmental goal set by the Swedish Environmental Protection Agency (SNV) is that the level
of Hg in I-kg pike (Hgpi), and in all types of fish consumed by man, must not exceed 0.5 mg/kg ww. This means
that most people would be able to eat fish from lakes without needing to consider health risks. The value Hgpi =
0.5 is used here as the critical value (Ecrit) to calculate geographical extent and duration of the Hg-problem in
Sweden.

1.4.1.2. Geographical perspective

Mercury concentrations in soil, water and lake sediments have increased by a factor of 4-7 in southern Sweden
and by a factor of 2-3 in the north of the country (Hellner et al., 1991; Lindqvist et al., 1991) due to increased Hg-
emissions during the lastcenlUry. Fig. 1.24 shows how the Hg-concentrations in the uppermost organic layer, the
mor layer, vary throughout Sweden. The distribution pattern can partiy be explained by known, historical,
domestic discharge sources. The largest point sources for Hg to air (1920-1987) are shown in fig. 1.25. A
comparison between figures 1.24 and 1.25 shows that the large discharges from the industries in Rtinnskiir,
SkuLSkiirlKorsnas and Bohus have caused considerable increases in the concentration of Hg in the mor layer and
that the discharges from point sources in DomsjtilKtipmanholmen, Ostrand, Str6msbruk and Skoghall have caused
a more diffuse influence.

The map in fig. 1.26 shows that the Hg-<'oncentrations decrease with the distance from a major discharge source
(R6nnskiir). This emission site has been chosen here to demonstrate the relationship between Hg in mor and
distance from a major point source in a situation where there is no (or only very limited) interference from other
point sources to blur the picture. For the R6nnskiir area, Hg-gradienLS in the mor layer may be observed in fig.
1.26 up to about 150 lan. The entire area of influence, as estimated using statistical methods (Nilsson et a\., 1989),
is up to 1000 lan, i.e., a very large impact area.

Levels in fish have increased fivefold in large parts of the country (fig. 1.27). The concentration of Hg in over
10.000 out of a total of 83,000 Swedish lakes is deemed to exceed the "blacklisting limit" of I mg Hg/kg wet
weight (ww) applying in Sweden (related to I-kg pike); about 40,000 lakes have fish (pike) with Hg-
concentrations higher than the guideline value of 0.5 (Nilsson et al., 1989). The levels are still rising in most
lakes.

34
Hg IG (1Ig/g dsl

m >0.:35
B 0.30-0.35
[j 0.25-0.30
('{ Q 0.20-0.25
V0 <0.20
o No observations

Fig. 1.24. Map of Sweden illustrating !be distribution of Hg (HgIG, llg Hg per g dry subslmlce =dry weight)
in ille mor layer. NOJe illat ille white parts in !be map lack data. From H1tkanson et al. (1990d).

1.4.1.3. Effect·load.sensitivity

Fig. 1.28 gives an ELS·model for mercury as a 3D·diagram. The operational effect variable Hgpi depends on how
ille pike has lived and what it has eaten during a long period before being caught (ille ecological half1ife for Hg in
l·kg pike is about 3 years; see Hirvenptiii et al., 1970; Olsson, 1977). TI,e Hg·concentration in pike is an integrated
value for !be entire environment of !be pike and its prey. The Hg-content in pike depends on ille Hg-Ioad supplied
to ille entire ecosystem, and not only to !be clump of reeds where ille pike was caugh~ as well as on ille biological,
chemical and physical conditions of ibis ecosysJem over a long period, since !bese conditions influence !be
distribution of ille Hg-Ioad on different carrier particles and !be bioavailability of ille Hg-Ioad. Hence, !be
biouptake of Hg in pike and its prey also depends on !be conditions in the entire lake. The pH of !be water is
irnpOrlmlt for !be binding of mercury to different types of carrier particles and for how Hg is distributed among
different Hg-forms. such as HgO, Hg+ and meJhyl-Hg.

35
Sufiljelma

.'Ostral,d 60 ton.s

(j
620
Malmo
15 tons

y (km)
'110 130 150 170 190

Fig, 1.25, Emissions of Hg to the annosphere (tons) from 1920 to 1987 from eleven major Swedish point
sources (black rings) and other sites. From Nilsson et al. (1989).

In the model in fig. 1.28, the dynamic ratio (DR =..JarealDm; area =lake area; Dm =mean depth) explains
statistically tlle greatest proportion (63%, or r2 = 0.63) of the variation in Hgpi among the 39 lakes (all data and a
derivation of the model are given in section 7.1; a dynamic version of the model is given in section 4). DR is
directly related to resuspension processes in lakes, the higher DR, the larger the lake andlor the smaller the mean
depth (see Hakanson and Jansson, 1983, for further information about DR and bottom dynamic processes in
lakes). The next most important factor, according to the stepwise regression, was the Hg-Ioad to me lake (HgS,
i.e., me median Hg-content of surficial sedinlent samples, 0-1 cm, reflecting me load to water and sediments for a
period corresponding to me mean age of mis sedinlentlayer). It is interesting to note tllat a sensitivity factor can,
in fact, be more inlportant man a load factor in explaining me variation in an effect variable. If pH and HgS are
pooled, 73% of me variation in Hgpi among these lakes can be statistically explained by this model. If one further
sensitivity factor is added, annual mean pH (pHI2), me r2-value increases to 0.80. The fourth model variable is
mean annual concentration of 10tal phosphorus (TPI2), which ralses r2 10 0.82. The lasllwo faclors are interesting

36
0.5

0.'

\ 0

u
. 0.3
0
0
0
0
0
,~
0 '» 0
~
0 0
0

-= 0
0
0 0
2 0 0 °oco
~ 0.2 0 0 0
J: 0 0 0 0
0 0 Q

0 0
if"
0
• 0 0
0
0

0
O. ,

Olstdnca from Rtlnnsknr (km)

Fig. 1.26. Data illusuating the decrease in Hg-concenuations in the mor layer with distance from the emission
at the Ronnskiir industry and the best-fit regression. From Nilsson et al. (1989).

since they demonsuates that the Hg-concenuation in fish is not only increased by a higher Hg-contamination
(given by HgS), but is also related to acidification, a case of synergistic effects, as well as with the third major
environmental problem in aquatic systems, eutrophication, a case of antagonistic effects. The greater the TP-
concenuation, the greater the production of algae, plankton and fish, and the greater the amount of biomass in the
lake. This implies that a given Hg-Ioad is spread over a larger biomass whereupon the Hg-coment in the biomass
becomes less ("biological dilution"; see Jansson et al., 1981; Hakanson and Peters, 1995). This is one explanation
why lake TP is important Another concerns internal correlations: If the bioproduction in a lake is increased, the
entire character of the ecosystem is changed.

The last factor in tbe model is tbe mean depth of the lake (Dm); which gives an r2 -value of 0.85. This is a high r2_
value for a model which covers such a wide range in lake characteristics, lake area from 0.12 to 307 km2 , mean
depth from 1.3 to 10.6 m, pH from 4.8 to 7.3 units and TP-concenuations from 6.9 to 4911g/l. The "empirically
based highest r 2 " for Hgpi is aboUl 0,86 (see section 7.1). This value is obtained by comparing two empirical
data-sets of Hgpi, which are supposed to give the same results. The observed difference has to do with
deficiencies in sampling and analytical procedures. This means that the given empirical model is close to tile
"higbest empirically based r2 ". But it is a regression model, and it does not give any time-dependencies for Hgpi.
The model is static. To obtain time variability, one needs a dynamic model (see section 4). It should be noted that
the given model can NOT be applied to lakes outside the given ranges, and it can neither be used for other types
of lakes, such as lakes from different climatological zones, nor for dams and reservoirs, where the Hg-
concenuations in fish are likely to be higher than the values predicted by tilis model.

37
Hgpi (mg/kg ww)

00 '1.25
III 1-1.21;
till 0.75-1
(J o 0.5-0.75
[3 <0.5
-" o No obse rvations

Fig. 1.27. The map illustrating the disUibution of Hgpi (i.e., Hg-content in I-kg pike) in Swedish lakes 1980 to
1988. The map is based on empirical data from 894 lakes and statistical calculations. This means that it is
likely, but not certain, that a given lake will have a Hgpi-value given by the contour line, but that there are
lakes within a certain interval with higher andlor lower Hgpi. Note that the white parts lack data. From
Hakanson et a1. (1990c).

1.4.1.3.1. Practical use of ELS-models in contexts of remediation

How can the model in fig. 1.28 be used in practice? This is discussed in greater detail in section 4. Fig. 1.29 (a
2D-version of fig. 1.28) exemplifies this here for a constant bioproduction (TP = 15 ~gfl, a typical value for a
mesotrophic lake; see table 1.5). Hgpi increases when the Hg-Ioad (HgS, the Hg-content of surficial sediments)
increases and the mean annual pH (pHI2, sensitivity) decreases. The figure also illustrates how this ELS-model
may be used in remedial contexts in a highly polluted (HgS = 750 mg Hg/kg ww; a natural background value for
HgS is about 100) and acid (pHI2 = 5.0) lake. On the basis of !his model, one can determine which measures
could be used to decrease the Hg-content in fish:

38
ELS-diagram. Mercury in lakes

=
For Area 10 kro2, Dm =10
m and pHI2 = 6,5

National environmental
goal for Sweden:
Hgpi< 0,5
O~O~.~----------

Nl,("j~ N
o ~-.::rC'"l
LOAD,HgS 16 SENSITIVITY, Total·P

Natural background value for HgS~ 150

MERCURY MODEL, see seaion 7: Effect variable for Hg in lakes:


lag(Hgpi)=Q,4 - 0,1 "pH12 - O,0041l9'TP12 - O,44*lag(DR) +
O,000397*HgS - O,IS*lag(Dm) Hm2i: Hg-conten1 in muscle in I-kg pike; in
=
(r2--o,8S, n 39: r:2=O,83 far Hgpi) mg Hglkg ww; at least 5 fish in the size range
from 0,5 to 1,5 kg per lake.
LEGEND:
Step 1. r:2=O,63. Sensitivity factar 1. Dynamic ratio • area resolution: entire lake for lakes smaller
(DR=-iArealDm); Area=lake area in kro2, Dm=mean depth in m. than ~ 300 km2
Step 2. r:2=O,73. Load factar. HgS=median Hg-content (mg Hglkg
dw) of surficial (0-1 em) sediments. • time resolution"" 3-4 years
Step 3. r:2=O,80. Sensitivitiy factor 2. pHI2=mean annual lake pH.
Step 4. r2=0,82. Sensitivity factor 3. Tatal-P (=TPI2)=mean All load and sensitivity variables should have
annual concentradon of total-P ()1g/l). the same area and time resolution as the
Step 5. r2=0.85. Sensitivity factor 4. Dm=mean depth (m). effect variable.

PRESUPPOSITIONS: Other effect variables for Hg in lakes:


Ranges of model variables
0,12 < Area < 307 None known., e«cept for Hg-content in
1,3 < Dm < 10.6 muscle for other species of fish used for
4.8 <pHI2< 7;3 human consumption.
6,9 < TP12 < 49

Fig. 1.28. ELS-model for mercury in lakes.

39
1. Lake liming; normal dnration ~ 8 years
Hgpi 2. Halting ditect emissions; recovery after ~ 10 years
(mg/kg ww) 3, 1+2.
3,00 4. Halting national atmospheric emissions; recovery after "'" 100 years
5.1+2+4 pH12
6. Natural conditions 5,0
2,50 Cwves for Area=l km2, Om
/5,5
= 10 m and TPl2=15 flg/l
.,,,,,' Common limit
_/~ .' 6,0 for lake liming in
2,00 ",,'" .... - Sweden
.,,,..-
;~
... -"
•• -
"..
,,6,5
.. ~ -0101------
,.'" .... -.... ~ /."".. ..,7,0
~.....~" .... " ..... .,.......... ......... 7,5
1,50 ~ .... --. -" / . ........... .."""
......... . .... ----:~ .... --.-
-!t'7';-:;...:';~~~~~O:::;::~'~-;'-;:',:",,=..:;;,;,:.::.-~:::::::::::.________ SWediSh blacklisting
1,00 >:"<oV' :-::.:-::;..:::::.~:-:..:-:..::__ '" limit; Hgpi=l

I~~~~~::::::===-________________________ Env,
0,50 :J goal;
Assumed Hgpi<05
namraJ pH12 Critical load at pHl2=7 5
0,00 +-j-+-+--+-+--+-+--+-+--+-+--+-t--i
o.,.,
~ HgS
(mglkg dw)

Assumed + Annospheric + Direct


natura! load load
load

Fig. 1.29. ELS-diagI1llIl based on the ELS-model for mercury in fish given in fig. 1.28. The figure also
indicates that practical guidelines, critical limits, natural background levels, environmenUll goals, etc. may be
defined from the model, and consequences of different remedial measures.

• The direct emissions of Hg to the lake could be halted (step 2 in the figure). But the Hg-content in pike will not
go down the day after direct emissions are halted. TIle lag-phase is rather long. Depending on the size of the lake,
it could take 10 years for the Hgpi-values to be adjusted to the new situation. The duration of the Hg-problem can
be simulated not with this static model but with the dynamic model given in section 4.

• The annospheric Hg-emissions could be reduced. This would cause a reduction in the annospheric Hg-
deposition and, in due course (centuries; see Hilkanson and Peters, 1995; and fig. 1.30), also in tile secondary Hg-
load from the catchment (step 4 in the figure).

• Sensitivity parameter I, pH, could be increased (e.g., by liming; steps I, 3 and 5 in the figure). The duration of
the remedy would depend on tile amount of lime added, the initial pH, the size of the lake and the lake water
retention time (see the lake liming model in section 4). For lakes of this lake type (glacial lakes), a liming would
often last about 4 years, and the ecological halflife time of Hg in I-kg pike is about 3 years. The duration of the
liming is regulated by the amount of lime added to the lake, lake morphometry (size and depul) and lake water
retention time.

40
• Sensitivity parameter 4, bioproduction (TP-concentration), could be increased (e.g., by fertilization, i.e., adding
of phosphorus; see Hakanson et aI., 1998 andlor sections 4 and 7). Naturally, the bioproduction must not be
increased indiscriminately to reduce the Hg-content in fish. Thalcould lead to a eutrophication problem!

Fig. 1.29 also illustrates that a natural Hgpi-value would be about 0.45 mglkg ww in this small, deep and
mesotrophic lake (area llan2 , mean depth 10 m, pHl2 = 7.5; TP12 = 15 fig/I; HgS =l()O mglkg dW). The
different remedies, from lake liming to halting of anthropogenic Hg-emissions, would bring Hgpi down from the
very high initial value of more than 1.5 to 0.45. The values in the figure illustrate the different routes to achieve
this end-point The critical load may also be defined from the ELS-model. This is the load were the effect
variable reaches the level given by the environmental goal. The critical load corresponds to a load of about HgS =
140 in the given lake, and is equal to the natural load (l00) in a relatively low-productive lake like this (TP = 15)
with a mean annual pH of 7. Such environmental goals and critical loads are of great imponance in water
management

This ELS-model may also be used to highlight an interesting, apparently paradoxical consequence of sewage
treatment. Assume that a city plans to build a treatment plant Before the treatment. the conditions in the lake are
assumed to be as follows:

• pH12 =7.4; no acidification


• TP-concentration = IS, tlle lake is meso trophic. All communes in Sweden must build purification plants because
the same rules apply to all communes (irrespective of the fact that the environmental conditions vary). The costs
for the plant is about 10 million US dollars.
• HgS = 450 (mglkg ww); the lake is contaminated by numerous diffuse emissions, e.g., from the urban area
(hospitals, dentists, small industries, etc.). The purtfication plant would reduce a significant part of those urban
emissions.

From measurements in the lake, and from the ELS-model (see fig. 1.28), the autllOrities would 1.'1l0W that the fish
in this lake has a very high Hg-coilcentration, Hgpi = 0.95, which is just below the "blacklisting" limit, but much
above the environmental goal orO.5. After the plant has been in operation (and new dynamic steady-state
conditions established), the following has happened: Direct emissions of phosphorus and, e.g., BOD (organic
substances causing a biological oxygen demand) have been reduced with 90%. The corresponding reduction of
nitrogen is 50-60%. The TP-concentration in the lake has been reduced from 15 to 10 figll. Lake bioproduction
bas decreased significantly, from 3.6 to 2.0 mg chlorophyll-a per m 3 (using the model in table 1.6). TIlis is
antiCipated and requested.

A lower production would also lead to a somewhat lower mean pH (see table 7.3, which shows that a positive
correlation exists between pH and TP-concentrations in lakes). The main reason for the reduction of 0.6 pH units
is that the high pH-values during the bioproductive season have decreased. The new mean pH is 6.8. TIle sewage
treatment has also lowered the direct Hg-emissions. Measurements reveal that 50% of the urban Hg is reduced
[360-0.5*(450-360)]. The new Hg-load is given by HgS = 320.

On tlle positive side, it may also be nOled that the Secchi deptll has improved. Thus, the lake looks cleaner and the
bioproduction has been reduced to about the preindustrtallevel. What about tlle Hg-content in fish?

41
Table 1.3. Predicted effects from the installation of a new sewage treaunent plant on a hypothetical lake, as
given by the as-model in fig. 1.28. Fish are blacklisted in Sweden when the mercury burden exceeds 1.0
mglkg ww.

Before After Note


Trophic level mesotrophic oligo-meso trophic
Chlorophyll-a (mg/m3) 3.6 2 from table 1.6

Total-P (TPI2, fig/I) 15 10


Lake pH (pHl2) 7.4 6.8
HgS (mg Hg/kg dw) 450 320

Hgpi (mg Hglkg ww) 0.95 1.02

Blacklisting NO YES

Table 1.3 summarises the results of this hypothetical but realistic example. TIle data show that after the treaunent,
Ihe aJJlhorilies would have to hlacklisl Ihe lake! The new Hgpi-value is 1.02. The main reason for this is that it
would never have happened had there been no aunospberic Hg-deposition from large scale emissions linked to the
use of fossil fuels and to industrial discharges. That part of the Hg-load is not controllable by the local authorities.
Another reason is that both bioproduction and pH were lowered, the sensitivity of the lake to Hg-<oontamination
was thereby increased.

1.4.1.4. Time perspective

Great effons have been made in Sweden to reduce the Hg-concentrations in lake fish (Hiikanson et al., 1991a). A
qualitative model has been used to produce curves (fig. 1.3'0) reflecting prevalent views about the relationships
between aunospheric deposition df mercury, mercury in pike and the reduction needed to achieve the
environmental goal (that all lakes should have fish with Hg-concentration below 0.5). Note, that there are no
values on the axis in fig. 1.30.

A dynamical model have been used (see section 7.2 for a description of the model) to put realistic values on the
axis of fig. 1.30 so that the duration of the mercury problem in Sweden can be address quantitatively.

An empirical mass-balance for the fluxes of mercury in a typical forest lake situated in the southern part of
Sweden is given in fig. 1.31. Atmospheric wet deposition of Hg is in the order of 20 g Hg/km2· yr, input via
tributaries 10-60 g/yr, sedimentation 10-40 g/yr and outflow 10-40 g/year. The Hg-evaporation is rather small, 1
g/km2'yr from land and 2-20 g/km2'yr from water. The uncertainties are very large for these fluxes. This simple
figure cannot show how very costly and time-consuming (Lindqvist et al., 1991) it is to produce the data in this
figure. And dry deposition may be considerable, but it is even more difficult to measure or predict, The main
message from the results given in fig. 1.31 is tl'at the Hg-concentrations in fisb in these Swedish lakes will not
decrease significantly until the Hg-load to the lakes decrease, and that will not happen until the aunospberic Hg-
deposition to the lakes and their catchments, i.e., the emissions causing the Hg-deposition, have been reduced to
sucb an extent that the deposition flux (200) is smaller than the input (l0-60).

42
Reduction=O%

---------------------eO%
Enviro~emal objective 0.5 mg Hg/kg
---------/80%

------~
Current Years Decades Centuries
situation
Fig. 1.30. The effects of reduction of atmospheric deposition of mercury on long-term patterns in mercury in
lake pike. A reduction of 80% is necessary to achieve the environmental objective of Swedish lakes set by the
National Swedish Environmental Protection Agency, thal all lakes should have fish with average mercury
burden below the 0.5 mg/kg wet weighL From Lindqvist et al. (1991).

Fig. 1.31. Fluxes of LOtaI mercury (in g/yr) calculated for a lake of 1 km 2 with a catchment area of 10 km 2 in
southern Sweden. Atmospheric deposition refers only LO wet deposition (not dry deposition). The amount of
Hg in the mor layer (i.e., the upper organic layer) of the soil is about 30,000 g in the entire catchment area.
Based on data from Lindqvist et al. (1991). From Hakanson and Peters (1995).

43
".0 <;I-ClJrve~ b~ed -l> ClJrvl" b~c4 on ~~umpt1on:s (drlv109 v.lu~) A.
en doh!
t..
">-., 1: S'wtdl~h omb~lo~ of H9 (toM:/~r)
.,
t..
0.
Z: Europo_n 'ml~ion.t of H9 rooehtng S'wtden (toM/Vr)

17.5
'"
::x:
"c:0
f-
,/' ' -----:'----22----
O.OI+~---_:_f;\....--s:.,-~':=~,=====~====::::,
1660 1965 2110 2235 Z365
o 125 250 375 500
Venr

B.
..... &I We.:sto disposal
a
~ ~ III Chloralkall !nd1l31ly
m~ 15,0 0 Cocl b.trning
E~ >-_ _ _--L = curve 1 1n A
ll.8 10.0
]-;;;,
o:r: 5,0
r o,o~~~~

Vear
I: 30000
2: 1.0 C.
3: 5.0

1: lSOOO
2:
3:
0.5
2.5
"------.3-_
I: Amount of H91n Qltc:hrnant (9)
3

Z: fl:sh e.oncentntion 0(1-19 ("'9/1::9)


'---
3: Concentntion 1n...,tor (nQll)
,:
2:
o
o
3:
OI~6~6:0----:-':96T.5~--------2~lr'0--------22~3-S--------2"36S
o 125 250 375 500
Vear
Fig. 1.32. A. Curves illustrating (1) calculated (before 1985-90) and assumed (after 1985-90) Swedish
emissions of Hg, and (2) assumed (= default) European emissions of Hg reaching Sweden. These two curves
are the driving load variables. From Hiikanson (1996c).
B. Quantities of mercury emitted from different sources (redrawn from Levander, 1989).
C. Model-calculated values of Hg in the catchment (in g), curve 1, Hg-concentration in I-kg pike (mg/kg ww),
curve 2, and Hg-concentration in lake water (ngll), curve 3. The model has been calibrated to yield realistic
data for the period from 1960 to the present (1985-90). The model is given in sec lion 7.

44
Tbe result of the model calculations is given in fig. 1.32C. The driving functions are curves 1 and 2 in fig. 1.32A
representing Swedish and European emissions, respectively. Fig. 1.32C shows that the Hg-concentration in I-kg
pike today in southern Sweden is about 1 mg/kg ww. This is also a realistic value for about 10,000 Swedish lakes.

Curve 2 in fig. 1.32C also gives the future time-course for the mercury burden of pike as predicted by this modeL
It will take a very long time: The values remain higher than the guideline value of 0.5 mg Hg/kg ww till about
year 2360, a very long time-perspective! Hence, T in PER is set to 10.

1.4.1.5. Cause and effect - a chemical theory

Since many biological, chemical and physical factors are linked to each other in aquatic ecosystems, and because
many factors are Important for the Hg-content in fish, it is not difficult to understand why it has been considered
hard to find simple mechanisms explaining why certain lakes have high Hg-content in fish and others have low.
Bj6rnberg et al. (1987) presented a chemical theory which explains why it is logical that pH would be so
important in these contexts. The approach, with its coupling between effect, load and regulatory factors, could
serve as model not only for Hg but for other substances as welL

The theory stresses some chemical reactions that take place outside the fisb and are linked to pH and Eh (redox
potential) of the lake water. The effect variable (E, tile Hg-contentin fish) is related to a Hg-load and a regulatory
factor; the dose or load tenn is given by the amount of Hg discbarged to the lake with primarily bumic matter
from the catchmen~ the regulatory factor is given by the activity of sulphide, selenide (and telluride), see fig.
1.33.

Tbe S 2- activity and its relation to pH and Eb is of special relevance. H2S dominates at low pH, S2- at high pH.
At the pH-values normally found in natural waters (4 - 9), the S2- activity is VERY low; 12 - 4 orders of
magnitude lower than tile toW sulphide concentration. H23 fanned through protonization of S2- will ventilate
away from acid waters, which will be continuously depleted of S2-. The activity of Hg2+ is primarily detennined
by the S2- activity. The solubility constant (Ks) of HgS(s) is 10-52 . This is an extremely low constant, and Implies
that practieally all Hg will be in the fonn of HgS(s) wben sulphide is available. The sulphide activity may be
reduced by a lowering of pH andlor an increase in Eb. Besides SUlphide, selenide and telluride display a similar
relationship to Hg (Ks = 10-58 and Ks =10-70 , respectively). The Hg2+ concentration in natural waters varies
within a wide range but is often in the order of about 5 ng/! (Hiikanson et al., 1990a). This is a high concentration
in tilis contexL Such high concentrations can only prevail when the S2- (andlor the Se2 -) activity is very low
(lower than 10-41 and 10-47, respectively). Thus, if waters have naturally high S2- andlor s.9-- activities, e.g.,
from sulphidic rocks in the drainage area, or if S2- andlor Se2- is added to the waters, then the Hg2+ activity will
be reduced most effectively. This means that it should be possible to treat natural waters with S2- as well as s.9--.
In practice, adding S2- would not be ideal since it would be effectively eliminated as H2S(g) or oxidised. This
would nOL to the same extent, be the case with Se.

Some results from Lake Oltertjarn, Sweden (see Bj6rnberg et al., 1988) can be used to illustrate how rapid and
effective the treatment of Se may be. The Se-concentration of the test lake was raised artificially from 0.4 to 2 - 4
!Ig SelL The increase was achieved by sodium selenite (Na2Se03) slowly leaching from a supporting material
(rubber). These data (in mg/kg fish muscle) demonstrate the quick response obtained between 1984 and 1985/86.

45
(f!!)
-
A +
[-factor
erg -c. 02::=;; C02
iI'
Bacteria

0- factor

Hum- Hg

C Hg- Hg+
• X - P04
X- [03

:.HqSe(s) ,
: K _10. 58
1-------1
I ______
L HgTe (s) ":

Fig. 1.33. Diagram illustrating the theory on the mechanisms regulating the relationship between the effect
variable (Hg-content in fish), the load factor (given here as Hum-Hg, i.e., the Hg associated to humic matter)
and the regulatory factor (the equilibrium between Hg2+ and regulatory ions (S2-, s,,2· and T,,2-) and the
linked relationships with pH and redox potential (Eh). The "sink" illustrates the most effective (Ks < 10-52)
trapping of HgS(s), HgSe(s) and HgTe(s). When, e.g., pH decreases, the S2- activity will decrease very rapidly.
This would increase the biologically available fraction, the Hg2+activity, and the possibilities for Hg2+ or
CH3-Hg+ to enter the fish. An increase in pH will cause the opposite result, i.e., an ultimate decrease in Hg-
content in fish, provided that there is sulphide andlor selenide available in the water. The equilibIium constants
for all other reactions are much higher than the constants of HgS, HgSe (and HgTe). Based on Bjomberg et al.
(1988).

PIKE PERCH
Hg Se Hg Se
Min. Mean Max. Min. Mean Max. Min. M~an Max. Min. Mean Max.
Before
n = 13 n=3 n= 21 n = 21
0.74 1.51 2.08 0.50 0.53 0.60 0.10 0.56 1.23 0.12 0.53 0.95
After
n= II n= 11 n=46 n =46
0.34 0.70 1.30 0.13 1.98 5.10 0.03 0.16 0.39 0.31 7.60 22.0

initially, one may expect that about 70 % of the Se added to the lakes would be attached to suspended particles
and thus have a limited effect on the Hg-content in fish. If a lake receives a cenain load of Hg primarily from high
molecular weight humic substances, what will happen? There are large numbers of bacteIia in the water which
use the carbon supplied by the organic material as a source of energy. During the bacterial decomposition, 02 is
consumed and C02 (and bacterial biomass) is created. Thus, tllere exists a link between the oxic conditions, the
redox conditions and tile activity of S2- and Se2- (i.e., the regulatory ions). When tile redox potential decreases,

46
the S2- activity increases and vice versa. But there is also chemical equilibrium between tile humic-Hg complexes
and Hg2+. The equilibrium constants of these reactions are in the order of 10- 10 to 10.20 (Smith and Martell,
1976), which is much higher tIlan tile Ks-values for HgS and HgSe. Similarly, there are equilibrium constants
between Hg2+ and metllyl-Hg or Hg complexes witll phosphates, carbonates, etc., which also ought to be in tIlis
order of magnitude. Assuming tIlat metllyl-Hg in organismslfish is connected to Hg2+ in water via a chain of
eqUilibrium reactions, this means that the presence of S2- or Se2 - will cause reduction of Hg levels or uptake in
biota. The exact nature of these equilibrium reactions (inorganic, organic or biologic) will be important primarily
for the rate of uptake or depletion of Hg. As experiments have shown, this rate is rapid enough to make addition
of selenium a practical remedy for high Hg levels in fish.

The theory does not address many important issues like (see Meili, 1991a,b; Watras and Huckabee, 1994; Hudson
et al., 1994): Where does the methylation of mercury takes place? Is Hg intake via food more important than Hg
intake via gills and skin? It should be stressed that the theory focuses on chemical reactions taking place outside
the cell that are linked to the pH of the water; which has been demonstrated by ample empirical data to be an
important sensitivity factor.

Another example concerning the interesting relationship between contamination, biouptake and conceniiation of
mercury in fish is given below. The data are from two lakes in tile River Kolbacksan water course, Lake Saxen
and Lake Bysjon (see Hakanson and Uhrberg, 1981). They lie close to one another. Lake Saxen is heavily
contaminated by effluent water from a mine (Saxberget), which emitted large quantities of, e.g., zinc in the form
of ZnS (from 1936 to the mid-1980s). This means tIlat large quantities of sulphur was also added to the lake.
Lake Bysjon is not influenced by any kaown major direct emissions. Bysjon is 5.1 km 2 and Saxen 0.8 km 2 One
hundred years ago both takes probably bad a pH around 7.5, a TP-concentration around 20 and a sediment
contamination around HgS = 100 (mg Hg/kg ww). This corresponds to a "natural", reference value for Hgpi of
about 0.6 for Lake Saxen (mg Hg/kg ww). Around 1980 the situation was as follows:

Lake n Hgpi year Hgpi year Area Dm pH TP HgS Hgpi-ratio


ernE: mod km 2 m 11g/1 mod/ernE
Bysjon 3 0.52 1979 0.66 1962 5.1 6.3 6.86 15 260 1.27
Saxen 11 0.20 1980 1.02 1976 0.8 3.8 6.97 22 540 5.1

It is a remarkable in itself that the empirical data for Hgpi (Hgpiemp = 0.20) for Lake Saxen are much lower than
the "naturrd" reference value of 0.6 but even more remarkable in light of the fact that the contamination has
increased by a factor of 540/100 = 5.4. The situation in Bysjon corresponds to the expected. TIle difference
between the measured value for Hgpi (= 0.52) and tile predicted value (0.66) is quite small and can be related to
the difference in the time periods (the empirical value was determined for pike caught in 1979, Ille model-
predicted value should be related to the mean age of tile sediments, which is 1962). What is the explanation to tilis
seemingly paradoxical result for Lake Saxen? The answer may lie in the given chemical theory (fig. 1.33). High
activities of S2- can be sustained in Lake Saxen due to the continuous contamination of sulphides for the mining
activity. The prevailing pH and Eh values in the lake will govern the distribution of S into various chemical forms,
including S2-, wbich bas a very high binding potential for Hg2+. So, a high level of S2- means a low level of
Hg2+, and everything related to this, like high mercury levels in fish. Note, however, IlJat one can NOT use
"sulphide lIeaunent" as a general remedial measure for high Hg-<=oncentrations in fish. This would be a very
expensive and inefficient roemod since most S2. will be oxidised in oxic lake water and since tile protonization
will remove much of me S2- into HS- or H2S.

47
It is evident that simple. general models of the type used in fig. 1.28 for mercury cannot account for special cases.
such as the conditions in Lake Saxen. But simple. general models are still very useful because they can predict the
normal. most likely result and give a reference value against which data from special cases can be quantified and
evaluated in a adequate scientific manner.

What remedial measures may then be used to lower mercury concentrations in fish (see also section 4)?

1.4.1.6. Remedial measures

The selenium method (or the "sulphide trealIDent") is. primarily. a theoretical approach to a very important
problem. but are there any ecologically. economically and practically applicable methods to reduce mercury
concentrations in lake fish? Fig. 1.34A summarizes the results for many tested remedies (see Hilkanson et al..
1990a). and some of these methods could also be used for other metals and radionuclides.

Intensive fishing can reduce the Hg-concentration in I+perch (= I year old perch) after 2 years by about 20-30%.
There is a causal link between Hg-concentrations in I+perch and Hg-concentrations in I-kg pike. since pike eats
perch. Allowing for the uncertainty in the prediction. one might expect Hg-concentrations in pike to decrease by
15-40%, but this will take some more years (see section 4) due to the difference in Hg-retention time for the twO
species. Intensive fishing can only reduce the Hg-concentration in fish substantially if the remedy is applied
rigorously. Then fish biomass has to be reduced more than 25%. Such an intensive fishing effort is very
demanding. especially in lakes larger than about 10 kril2 The effect of intensive fishing would last about 10 years,
because it generally takes about 3-5 years before the Hg-concentrations in pike decrease to minimum values and
presumably about 3-5 years for the concentrations to increase again. Intensive fIshing is not a cost-effective
remedy for Hg, if anglers or other workers are paid current salaries. However, it may be a suitable measure for
fIshing clubs whose members enjoy fIshing and wish to help to improve the Hg-situation in a lake. The results of
intensive fIshing could be improved in acidic lakes if the remedy is combined with lake liming andlor wetland
liming to change the water chemistry (pH. hardness. etc.)

Drainage area liming with dolomite was carried out in four catchment areas. This is an expensive remedy. It
affects the entire ecosystem. the forest. soil chemistry. life in streams and lakes, and possibly also fIsh Hg-
concentrations for a long time. After 2 years. liming of the dralnage area increased lake pH by about 8% on
average. Tile basic data for the dolomite trealIDent are uncertain and the lakes need to be fallowed for several
more years.

Selenium treatment involved a dOSing problem Ulat must be solved before this method can be recommended for
more general use. Because Se is itself toxic, Se-concentrations in water must not rise above about 1-2 fig Sell and
the Se-concentrations in fIsh should never be higher than about 2-3 mg Seikg ww. If the selenium load can be
controlled this well, the method holds promise. particularly for large and severely Hg-contaminated lakes.
However, there are questions concenting the effects of selenium on oUler species of animals than pike and plants
that must be elucidated before the method can he recommended more generally. In particular. it is remarkable
that the reproduction of perch decreased drastically in rouror the eleven lakes treated with selenium (all
used the Botiden method; using selenium sealed in rubber rubes placed in nets in the lake; see Paulsson and
Lundbergh, 1989). The data indicate that this reduction is probably due to the selenium treatment, even though the
mechanisms involved are not known. The experience with the selenium method suggests that the dosing problem
may be reduced if the selenium nets are placed near tributaries. where tile water rurnover is fast, and if the nets are
in the lakes during the spring, summer and early autumn. but not under the ice.

48
Measuresvs Hg in l+perch 1989
100 A. 1 ;oU !>rror "or.
90,--------------------____=-_________ ~
80,------------ - - f - - - - - f
<!'
70i-----------------~--------f
.s
60--.~----------------~~~\~----4
" 50
III

e ;-:~ -li ~~---::----

~l -CI"'I!~ I!" I "1 gHq.~~


u
o"

=--·"'II",!I""'II."'i!i"'=-1
j
O~~~~~~~~~~~~~~LJ
U. WLL DAL Se IF
n=26 n=23 n=4 n=7 n=B

~
(n=ll)

Lake liming
100
90..E· ' 1 SD Error
80j----------------------------1
<l1 70 - - - - - - - - - - - , - - -------------f
~ 60,----.--------1---------------------
!'J SO,--+
[ ______ 1__ - - - j - - - - - - -..TT---
b 401-r.~J~···~---t_-----I----~~~I~t
25 30 - ...... ··-:I-...",~~l----I------- . .:1-
~~ -- .:-:::~ ?)I,--t~,:=i<b:::'9"+--I-:·::t
O~~~--J~~~~~___~~~
PR&SR M FER pH&POT
n-9 n=S no3 n=6
Fig. 1.34. A. Comparison of how different remedies (LL = lake liming, WLL = welland liming, DAL =
drainage area liming, Se = selenium treaunem plus liming and IF = intensive fishing) influenced the Hg-
concentration in I+perch. Values for 1989 are expressed reiative to values from before treatment (1986).
B. Comparison of the influence of different types of lime used in lake liming (pR = primary bedrock lime, SR
= sedimentary rock lime, M =mixed lime, FER =fertilization, pH&POT = liming to higher pH plus potash) on
Ibe Hg-concentration in 1+perch (Hgpe) following remediation. Values for 1989 are expressed relative to
values from before treatment. From Hiikanson and Peters (1995) based on Hiikanson et al. (1990a).

Lake liming. There is very good agreement between the effect of liming on mean pH of Ibe lake and on Hg-
concentration in 1+perch (Hgpe in mg/kg ww). When liming is done welL the Hg-concentration in perch fry is
reduced by about 30% in two years. The concentrations in pike may therefore decrease by 15-40% after two or
more years. The type of lime did not affect the reSUlts, for there is no difference in the reduction of Hgpe-values
among bedrock lime, sedimentary rock lime and "mixed" lime. Since lake liming is conducted on a large scale in
Sweden to minimize acidification, it is important and interesting that it also reduces the Hg-concentration in fish.

Wetland liming. Similar reductions in the Hg-burdens of perch fry were obtained for lakes treated with wetland
and lake liming, and again there were no significant differences among different types of lime. Consequently, the
most important aspect of the remedy appears to be how efficiently Ibe liming is carried out, i.e., how much change
is induced in pH and in all the water chemical variables correlated with pH. The Hg-concentration in perch fry can
be reduced by about 30% in two years, so again, tlle concentrations in pike may, in due course, be assumed to
decrease by 15-40%. Wetland liming does not appear to influence lake colour or the Hg-transport from soil to
water to any extenl

49
Fertilizmion. There is nothing in the data from the "Liming-mercury-cesium" project to suggest that fertilization
can reduce the Hg-concentration in perch, but the fertilization program in this study was not effective, since there
were only small changes in total-P concentrations. Fertilization may still be able to ameliorate Hg-contamination,
as discussed in other studies (Hiikanson, 1980a). This will also be discussed in section 4.

1.4.1.7. Summary - mercury

Utilizing the PER-criteria, one can note that:

• Mercury contamination does NOT seem to influence reproduction Or biomasses of key functional organisms in
aquatic ecosystems. The generally used effect variable is mercury concentration in predatory fish (I-kg pike).
Lacking convincing scientific evidence concerning ecosystem effects for mercury in aquatic ecosystems, and
using the precautionury principle, the E-value is' set to 3, wltich means "likely but low ecosystems effects using
statistical (i.e., probabilistic) calculations". The "critical" E-value (Ecrit) is set at the environmental goal, Le., 0.5
mg Hg/kg ww.
• Mercury contamination has, like acidification, a very wide geographical spread. The A-value is set to 8 since
about 40,000 lakes out of 83,000 have Hgpi-values higher than EcriL
• The Hg-contamination (i.e., Ecrit > 0.5) will likely continue for several centuries. The T-value is set to 10.
• PER is high (= 3*8*10 = 240) for the mercury contamination of Swedish lakes. The operational effect variable
Hgpi is, however, quite different from tile effect variables for acidification (such as extinction of roach). So,
mercury contamination is a smaller environmental problem for freshwater ecosystems than acidification (PER =
240 as compared to 700). Hence, to save Swedish freshwater ecosystems, more effort (time, remedies, research,
money, etc.) should be directed to minimize the acidification problem than the mercury problem.

1.4.2. Radiocesium

The nuclear accident at Chemobyl. togetiler with unfavourable wind/weather conditions in Scandinavia, resulted
in large parts of Sweden being exposed to a major fallout of radionuclides in late April-early May, 1986 (Moberg,
1991). The alpine areas of Sweden as well as large parts of the coastal and inner areas of the county of Norrland
were exposed to Ule highest fallout (max. about 100 and mean = 25 kEq/m2) of cesium-137 (fig. l.3SA).

1.4.2.1. Effect variables

".... if man is adequately protected Ulen oUler living Ulings are also likely to be sufficiently protected". This
SlJltement from Ule International Commission on Radiological Protection (lCRP, 1977) is based on extensive
reviews on me effects of ionising radiation, including radiocesium. Today there are many ongoing studies
witilin the framework of international organisations (such as IAEA, the Interrtntional Atomic Energy Agency,
Vienna) and several EU-projects (such as MOIRA, see Appelgren et al., 1996; Monte et al., 1997) to structure
the existing knowledge and do research in the field concerning radiological effects to ecosystems. Fig. 1.36
shows the relative tolerance of different groups of aquatic organisms to the radiation dose required to kill 50%
of exposed individual in a given period of time. Mammals are most sensitive and hence the target group in tilis
context.

50
Cspe (kBq/kg ww)

m 15-20
o HHS
[J 6-10
03-6
[) 1.5-3
0<1.5

Fig. 1.35. A (left). Estimated fallout of radiocesium (Cs-137 in kBq/m2) in Sweden after the Cbernobyl
accident (AprilfMay 1986). From Persson et aI. (1987).
B (right). A map illustrating the disuibution of lakes witil different radiocesium concentrations in 100-g perch
during fall 1987. NOle that this map is based on empirical data from 644 lakes and statistical calculations. This
means that most lakes (but not all) are likely to bave fish with cesium concentration given by the contour line
values. From Hiikanson et aI. (1992).

Tbe effects on aquatic organisms from ionising radiation concern both individual and population levels and
the variables of interest are, e.g., mortality rate, fertility rate and mutation rate (see Whicker and Schultz, 1982;
Jimenez and Gallego, 1998). There are major differences between diJferent organisms in sensitivity to ionising
radiation, and there are major differences for one and the same organisms living in different environments. This
is reflected in the wide intervals given in fig. 1.36. The recommended dose rate limits (from lAEA, 1988) are:

• below 1 mGy/hr (Gy = grey, see below for definition), no population effects,
• 1-10 mGy/hr, decrease in reproduction,
• above 10 mGy/hr, increase in mortality.

51
II Protozoa I

I Algae

I Mollusca I

I Crustacea

I Fishes I

IManl

1 10 100 1000 10000

Gray

Fig. 1.36. The relative tolerance of different aquatic organisms to the radiation dose required to kill 50% of
exposed individual in a given period of time (from USNRC, 1978).

The National Council on Radiation Protection and Measurements of the United States (see USNRC, 1978) bas
taken a somewhat different approach and stated that 10 mGylhr is sufficient to the protection of aquatic
organisms.

Ionising radiation can result in damages at cellular and whole-body levels. Molecular changes may be induced
in the DNA. When effects sucb as chromosomal alterations do not result in cell death, they can be transmitted
to subsequent generations. The response to the radiation can be modified by several factors, such as the type of
radiation (different types of radiation, such as a. ~,'Y, have different biological effectiveness) and the dose rate
(a higher total dose is required at low rates to produce the same effect than at high dose rates). The existence of
a threshold of dose rates over which effects are produced has been the object of a long debate. DNA repair,
which operates on a molecular level, may have a particular significance at low dose rates.

Environmental factors can also influence responses to radiation, e.g., in mammals. Radioresistance increases
under anoxic conditions. Salinity tend to increase radiosensitivity due to the metabolic stresses they demand on
the organism. Radiation response also depends on the stage of the cycle the cell is in when the exposition
occurs. In general, the radiosensitivity of cells is directly proportional to their rate of division and inversely
proportional to their degree of differentiation (lAEA, 1988).

There are also indirect effects of radiation, e.g., more resistant organisms can eam competitive advantage over
others less resistant groups. The difficulties of getting meaningful results aboUl radiation effects on natural
populations in their natural environments have limited the available information on many species and
communities, and very little information bas been obtained about the interactive effects on radiation with other
agents or stresses, and the long-term effects of chronic low-level radiation.

The measurement Or calculation of the dose rates experienced by organisms is necessary in order to assess the
impact of the release of radionuclides to the environmenL The fundamental unit of radiation absorbed dose is

52
the rad, defined as: 1 rad = 100 erg (absorbed)/l g material. In the International System of Units, the absorbed
dose unit is the grey (Gy), which is equal to 100 rad or 1 JJkg. The fundamental unit of radiation exposure,
which only concerns X- or y-radiation, is the roentgen (R), which is defined as the production by X- or gamma
rays 1 electrostatic unit of charge of either sign per cubic centimetre of dry air at 0 degree centigrade and 760
ram mercury. 1 R =87 erg/g =0.87 rad in air.
Aquatic organisms will receive external exposure from radionuclides in water and sediments and internal
exposure from nuclides accumulated in their tissues. The calculation of such doses is the goal of several
dosimetric models (IAEA, 1988), which use as input the amount of radionuclides in the water.

A problem arises when considering tile contribution of a., ~ and y radiations. It is customary the use the same
radiation weighting factors employed for humans, i.e., 20 for a.-particles and 1 for X-rays, y-rays and electrons.

The external dose rate received by fish in water may be calculated accordingly (from IAEA, 1992) using Cs-
137 and Sr-90 as type substances:

External exposure of fish from water (mGy/yr per BqlI)


~+ y radiation y radiation
Cs-137 0.0032 0.0027
Sr-90 0.0027 o
The external dose rate of fish (De(S) in mGy/yr) from sediments may be estimated as (IAU, 1992):

De(S) = 0.00252*C(S)*E (1.1)

where
C(S) = the sediment concentration in BqJkg W\V;
E =the total energy of tile decay (MeV). as given by:
~ radiation y radiation
Cs-137 0.187 0.56
Sr-90 1.1

The internal dose is given by:

(1.2)

where
C(i) =tile average concentration of tile radiolluclide in tile tissue in BqJkg ww;
Ed = tile radionuclide decay energy (MeV) effectively deposited in the organism given by (from USNRC.
1978):

Energy deposited per disintegration (MeV)


Fish Invertebrates
Cs-137 0.5 0.27
Sr-90 1.14 1.14

53
The factor 1.6*10- 13 has the units of Gy/yr per MeV and equals 1.6*10-6 ergslMeV times 1*10-7 Joule/erg.

Lacking well-tested target effect variables related to the reproduction or survival of key functional organisms in
real aquatic ecosystems, the concentration of radiocesium in fish muscle consumed by man will be used in the
following PER-analysis as a simple operational effect variable, and the guideline value for commercial marketing
of fish recommended by the Swedish National Food Administration of 1500 Bqlkg ww will be used as Eerit.

1.4.2.2. Geographical perspective

The radioactive fallout led to lake fish in about 14,000 Jakes (daring fall 1987) in the afflicted areas receiving
concentrations above the guideline limit of 1500 kBq/ kg ww (see Andersson et aI., 1990; fig. l.35B). These
consequences of the Chernobyl accident (together with a very intensive mass media debate) have implied a major
economic setback for those parts of the tourist industry based on'sport fishing, but have also caused problems for,
e.g., fishery management and local population who use lake fish as part of their basic diet (Bengtsson and
Hansson, 1990).

1.4.2.3. Effect-load-sensitivity

An empirical ELS-model (r2 = 0.90) for radiocesium in Jakes is illustrated in fig. 1.37. The guideline value for
Sweden, 1.5 kBq per kg ww, where the authorities place restrictions for human consumption, is marked with an
arrow. It should be noted that the model in fig. 1.37 only applies for a given year after the Chernobyl accident.
The most important senSitivity factor in this model is Jake conductivity - the more ions similar to Cs. like K, Ca,
Na and Mg, the higher the conductivity and the lower the uptake of Cs-137, a case of "chemical dilution" (Black,
1957; Fleishman, 1963; Carlsson, 1978). For a "peak emission" like this, the biouptake of Cs- 137, and the Cs-
concentration in fish, is lower in lakes wiill fast water turnover than in lakes with slow water turnover - a case of
normal "water dilution". The opposite is valid for Hg, w~ch is supplied to the lake "continuously". In that case,
the biouptake increases with incr~ased runoff of Hg and water from the catchment, i.e., wben the Hg-Ioad from
the runoff is renewed many times per time unit (Meili, 1991a, b). An increase in total-P, i.e., in lake
bioproduction, causes "biological dilution" for both Hg and Cs-137 (Hlikanson and Peters, 1995). These
relationships are quantified in the ELS-model in fig. 1.37 and during the model derivation the importance of the
model variables were ranked in relation to other factors that could also be assumed to influence the given
operational effect variable.

Cs-137 seems to bave a considerable particle affinity (Broberg and Andersson, 1989; Riise et aI., 1990; Madruga
and Cremers, 1997; Konitzer and Meili, 1997; Konoplev et aI., 1997). This is very important and means that the
behaviour in lakes is strongly influenced by the chemical and physical properties of the carrier particles. This is
certainly the case for Hg (Lindqvist et aI., 1991). In addition, the retention in a lake will be governed by the
turnover time of the lake water; the longer the water turnover time, i.e., the larger the proportion of the initial
cesium load will be retained wiUlin the lake (mosUy in U,e sediments).

54
Empirical ELS-model - cesium in lakes

EFFECf VARIABLES (examples):


SENSITIVITY V ARlABLES (examples):
- Cs-137 in fish for human consumption:
- Cs-137 in pike; -lake water retention time (Tw, years)
area constant=entire lake (if a<1O \an A 2); - lake water chemistry of the "cluster"
ecological halflift? 10 years variables, i.e .• cond, alk. CaM"g or pH; mean
-Cs-137 in perch fry (I+perch); values for the entire lake for a time equal to the
area constant=entire lake; time constant..
ecological halflif"", 1 year - lake morphometry, especially "cluster"
parameters linked to resuspension. i.c .• mean
.. No known threats to the aquatic ecosystem depth, dynamic ratio and BET (=area of
erosion and transponation).
LOAD VARIABLES (examples):

- Fallout of Cs-137 ("Primary load) For Tw=l year


- Secondary load (=discharge from catchment)
.. Mean lake concentrations ofCs~137 as
determined from: EFFECT, Cspi87
- water samples
- sediment trap material or
- surficial sediment data

1500 Bq/kg ww - ----t>


Swedish guideline

LOAD, Cssoil
SENSITIVITY, cood12
Model (HAkanson. 1991):
log(Cspi 87)=0.783 'I og(Cssoil)-O 51 7 '..Jcon d 12-Kl.23 3'..JTw -Kl581
(r A 2=O.90; n=41)

Effect variable:
- Cspi87; Cs-137 in pike caught in spring 1987 in Bq/kg ww

Load variable:
- Cssoil; mean fallout on lake and catchment in Bq/m'2; range 2500-70,000

Sensitivity variables:
• cond12; mean annual lake weater conductivity in mSjm; range 1.6-8.2. and
.. Tw; theoretical lake water retention time in years; range for 0.02-2.9 years.

Fig. 1.37. ELS-diagram for cesium in pike according to an empirical model, which gives Cspi87 (cesium in
pike in spring 1987) as a function of Cssoil (fallout of radiocesium after Chernobyl), lake water conductivity
and the theoretical water retention time (for Tw =1 years).

55
Many of the processes controlling the flow and biological uptake of Hg and Cs-137 in lakes are controlled by
hydrological and morphometrical factors (see the model in section 3.1), which cannot be influenced by remedies
changing the water chemistry, e.g., liming and potash treatment (see section 4). On tile other hand, other processes
are clearly linked to the water chemical conditions of the lakes. Remedies that could speed-up tile recovery (=
reduction of the elements in fish) must therefore aim at either reducing the uptake in biota by blocking the
biouptake, increase bioproduction, andlor include specific elements, like potash (K) that can be taken up in fish in
about the same way as Cs-137 (Black, 1957; Fleishman, 1963; Carlsson, 1978), or Se, which would effectively
bind Hg (Rudd and Turner, 1983; Rudd et al., 1983; Bjornberg et al., 1988; Paulsson and Lundbergh, 1989).

The water turnover time (Tw) is kept constant (1 year) in the diagram in fig. 1.37. The conductivity varies in steps
of 1.6 mS/m, Cssoil (fallout of radiocesiurn) is given on th~ positive x-axis and the operational effect variable
(Cspi87) on the positive y-axis. From this diagram, it is easily seen how the variables are related to each other.

1.4.2.4. Time perspective

Fig. 1.38A shows tile development of Cs-concentrations in pike, 80 g perch and I +perch (Le., fish 1 but not 2
years of age) in lakes with a fallout of 50,000 Bq/m2 , a water retention time of one year and a conductivity of 4
mS/m. It can be noted that in a lake of this type, the small 1+perch has steadily decreasing concentrations of Cs-
137, which in October 1988 bas not reacbed 1500 Bq/kg. For pike, the Cs-concentrations continue to increase
until about 1990 (see fig. lAO). Fig. 1.38B shows tile corresponding temporal development for trout and char. To
predict the time-dependent recovery, one must generally use dynamic models, i.e., time-dependent models based
on differential equations (see section 3.1).

The intention here is to provide both a geographical and a temporal piclllre of the development for the operational
effect variable in the PER-analysis, the Cs-concentration in fish. Fig. 1.39 gives maps for this for the years 1990,
2000 and 2010. The base for these maps is made up of empirical values of conductivity (from Monitor, 1986) and
tile fallout (from fig. 1.35A), together with a dynamic model (Hiikanson, 1991), which gives the calculated Cs-
concentrations in pike for these three years. Fig. 1.39 shows the results of a simulation where the water turnover
time has been held constant at one year, which is typical for Swedish lakes (Hiikanson and Peters, 1995). The
fallout varies between 2000 and 70,000 Bq/m2 and tile conductivity berween 1.5 and 50 mS/m. The maps are
intended to give the mean concentrations for different regions. The mean concentration in individual lakes or the
Cs-concentration in individual pike, may naturally diverge considerably from the values given in the figure.

As can be seen from fig. 1.39, it is the areas that received a large fallout that have, and in the fulllre will bave, the
highest Cs-concentrations in pike. In 1990 there were about 7000 lakes with pike having Cs-concentrations in
excess of the Swedisb guideline value of 1500 Bqlkg ww.

It must be emphasized that the dynamic model utilized to produce the maps in fig. 1.39 was developed shortly
after tile Chernobyl fallout in 1986. It predicts a decrease in Cs-concentrations in pike that is about 10 years toO
fast (fig. lAO). The new dynamic model given in section 3 accounts in a much better way for the processes
regulating the recovery time, like the secondary load from the catchment, tile distribution berween dissolved and
particulate pbases for radiocesium, sedimentation, internal loading and the outflow of radiocesium. Many of the
older models predicted unreallistically fast recoveries, whicb is demonstrated in fig. 1.40 wbere the new model
(section 3.1) is compared to an old model (from Hiikanson, 1991) and to empirical data. Even tilOugh the old

56
16000

... [-
I "000

~
12000
C'
e
"'"XJO
'"•
C
E 6000
<-
".;
~
6000
U
4000
o Ow

2000

o
Apr Ocl ft.pr Oel A.pr Oct
\986 \987 \968

12000

.
i
'"0'
~
10000

8000
e
.<:

C 6000
.!l
<-
"
~
.000


U
2000

1986 1987 1988

Fig. 1.38. Upper: The temporal development of radiocesium in large perch (> 80 g), small perch « 12 g) and
pike when fallout is 50 kBq/m2 , theoretical water retention time I year and water conductivity 4 mS/m.
Lower: The same for char and brown trout in Swedish mountain lakes. From Hakanson et al. (1992).

model predicts the initial empirical concentrations rather well the "tail" of the curve is not realistic. This motivates
wby aboutIO years should be added to the time scales given in fig_ 1.39. The model calculations given in this
figure indicate, e.g., that in year 2010, the Cspi-values will. on average, be in the range 100-300 Bq/kg ww within
the most exposed areas in Sweden. The maximum Cs-concentrations in pike in this year should be below 1500
Bq/kg ww. In the year 2020, almost all lakes should have Cspi-values lower than 20 Bq/kg ww.

From recovery models andlor from empirical time-series of dala, one can determine values for the ecological
halflives for cesium in fish after the Chernobyl event. Tbe ecological balflife reflects the actual decrease in
contaminant concentration for fish in real ecosystems, while the biological halflife reflects the decrease in
individual fish if they are place in uncontaminatedenvironrnents and the decrease is regulated by melabolic
activities. The new dynamic model (section 3.1) gives an initial ecological halflife, from month 30 when the peak
value of 28,000 Bq/kg ww is attained to month 70 when half that value is attained, of about 40 months for pike in
the Finnish lake, Iso Valkjarvi, which has a long theoretical water retention time of 3 years (fig. 1.40). Note that
this initial ecological halflife has been determined relative to the year when the maximal Cs-concentration

57
2000 2020
2010
1990 2010
2000
Cs·137 tn pike
(Bq/kg ww) Cs·137 in pllte Cs·137 In pIke
~ :lOO~OOO (Bq/kg ww) (Bq/kg ww)
ri 1$'JO~3000 IB 101)-300 o '~20
o :100·1$00 Q 60-100

0< 300 0<60

Fig. 1.39. Recovery prognosis for radiocesium in pike in Swedish lakes according to a dynamic model
presented by Hakanson (1991). Note that this model, which was developed a few years after the Chernobyl
accident in 1986, predicl.5 a recovery which is too fasL To get more realistic predictions one must add 10 years
to the initial values.

appears in this species of fish for this particular lake. The halflife is longer than 170 months if one startS with the
value of 14,000 (see fig. lAO). This is a typical pattern. Initial halflives for radiocesium in water and biota after
the Chernobyl fallout are generally much shorter than subsequent halflives.

1.4.2.5. Summary ~ radiocesium

• Contamination by radiocesium does NOT seem to influence reproduction or biomasses of key functional
organisms in aquatic ecosystems at these fallout levels « 100 kBq/m 2). However, a caveat in this context is that it
may take very long for potential effecl.5 to develop. A generally used operational effect variable is radiocesium
concentration in predatory fish reflecting the (foodstuff) aquatic pathway of radiation exposure to man. The E-
value for radiocesium in lakes is set to 2, which means "unlikely ecosystems effecl.5 using sUltistical, probabilistic
methods".
• The con!a1Ilination of radiocesium after the Chernobyl accident has a wide geographical spread. The A-value for
Swedish lakes is set to 6. This means that lakes in "many regions" have fish with radiocesium level higher than
EcriL set at the guideline level of 1500 Bq per kg ww.
• The contamination of Swedish lakes will likely continue for several decades, until about 2030 for the lakes with
the slowest recovery. The T-value related to the given Ecrit-value of 1500 Bq/kg ww is set to 6 ("more than 20
years, less than 40 years").
• PER is 72 for this contamination, but tile effect variable, Cs-concentration in fish consumed by man, is different
from the effect variables for acidification (which are based on real reproduction damages to key functional
organisms).

58
Month 30 Mont h 70
L.ak" Iso Valkjor-vl
Max 30000 T = 3 yr
C,Mg" 0.1 meq/l
28000
~

:.
:.
-0>

'"c r
'"
~

~
Empi rical data

'" 15000
'0. 14000 Ne ... model
.::<= 240

--...
0

<=
7000
~
0
<= Old model
0
0
I
n
U
0
61 120.5 180 240
Month

Fig. lAO. Illustration of the recovery time and the concept of ecological halflife using data from Lake Iso
Valkjtirvi, Finland, which is a small (area = 0.042 kIn 2), oligotrophic, softwater lake (hardness = CaMg = 0.1
meqll) with a long theoretical water retention time (3 years). One curve shows the situation as predicted by the
new model for radiocesium (see section 3) and one curve shows the values as given by an old dynamic model
(from Hiikanson, 1991). The empirical data are given by the bars. The max.-value for Cspi is attained month 30
(month 1 is January 1986) and the initial ecological halflife for cesium in pike in this lake is 40 months. The
subsequent recovery is much slower.

1.5. Chlorinated organics

Fig. 1.41 gives one example how io structure the very complex group of chlorinated organic material. There are
several levels here, where AOX stands for adsorbed organically bound halogen, TOCl for total chlorinated
organic material, EOCI for extractable organically bound chlorine, EPOCI for extractable persistent organically
bound chlorine, etc. (see table 1.4). The focus in this section is on emissions of chlorinated substances from paper
and pulp mills. This means tilat many aspects of halogenated substances in aquatic ecosystems are not treated (see
SOdergren, 1992).

1.5.1. BaCkground

Discharges of chlorinated organic materials from forest industries have attracted a great deal of attention,
especially during the late 1980s in Sweden and Finland. Most examples here deal with data and results from that
period. This is a complex group of substances (fig. 1.42; see also table 1.4) which includes some of the most well-
known and discussed environmental toxins, such as dioxins and probably a considerable number of substances
about which little is known today as regards their properties andenvironmental effects. There are no validated
ELS-models for substances belonging to this group. Lacking such information there is ample space for
statements and speculations about which ecological effects occur as a result of given substances and what can be
linked to other causes.

59
Nutrients I Acidifyin9 Toxic
Eutrophiea.tion S\Jbshncil!'S' substanetS

Chlorinated
org.nics
I
~EJ I

Radioisotope:

T
TOCI

... r- ... ~ ... .,. ... r-


Volati1~ Lipophilic Water solub. High moleeu-
substances ~ substances $UbS"tancfils 1..- sob", •

... ,.. .-:' P-

Other
EOCI
compounds Other known
,.. ~
subshncill's

T , "
Unknown
Pers. subS't.
EPOCI
Non persist.
eompunds
~

"
,
to. substances

V
T+ .. to. to. ...
I
r I
" I
" I
I
"
Dioxins IIFunrws PCB DDT/etc Toxafen Chlorophenol~ HCB

-".
-j,'
• Ii"
other known Unknown Other enlor Trichloro-
S\Jbstances subshnces pheno1s l 9Uaiacol

Fig. 1.41. A differentiation scheme for chlorinated organics in marine areas. From Hakanson (1990).

60
Table 104. Some well-known chlorinated organics.

IQQ: total organically bound chlorine


AQX:. adsorbed organically bound halogens
EQQ: extractable organically bound cblorine
E£QCl: extractable (acid-)persistent organically bound chlorine
dioxins: PCDD or polychlorinated dibenso dioxins; and furanes; there are many dioxins and furanes, of which,
"the diny dozen", are considered of special interest in ecotoxicology
~: polychlorinated biphenyls; lipophilic substances used, e.g., in oils; certain forms, like planar-PCBs are

considered to be responsible for the sterility of Baltic seals .


DDT: dichloro-diphenylchloro-methylmethane; this group includes, e.g., lindane, aldrine, dieldrine and DDE,
all well-I."11own from R. Carson's book "Silent spring"
HCB: hexachlorobensene
HCli: hexachlorocyclohexane.

fi.

O-C-H
l'
1. Simple aliphatical hydrocarbons: ~,
e.g., chloroform (a)
carbontetachloride (b) j.~
meiliylchloroform (c) PCW

vinylchloride (d)
2. Aromatical compunds:
k. 6
r
e.g., chlorophenols (e) c. C1 H
hexachlorobesen (HCB), (I)
chlorateluen (g)
3. PolycyclicaJ aromatical compunds:
:yt-« I I
c
I
e.g.. polychlorinated biphenyls (PCB), (h) c-
polYChlorinated naphtalenes (PCN), (I) I
c
polychlorinated dibenso-p-dioxins (PCDD), (j)
4. Complex organical compounds:
1~:
e.g., lignins (k)
e. 0<

humic acids (I)

f.
t o
,
'"
0
0
I
R
...... ~

CI
Ovlwc\
Of8A.c1
b
g. H
~-CI
, H

h.
~'
~

Fig. 1.42. Different carbohydrates of general interest in aquatic ecotoxicology. Note that all Cl-atoms are not
added to all molecules, that the smtcture of lignins only gives a small part of this very large molecular complex
and that the humic acid only gives one example of what humic acids may look like. From H1tkanson et al.
(l988b).

61
:",.;'. ",: ::': iox;'pj,i:~;"": .
A.
:.Chl~r;'p.llenols ::,:"OOTs:
·:i>COOs:: B. Differentiation ofEOCL

!liJJtJ/~tln@tjj
@Jfl1ga«I1d.JI@
~/rt:
dlwl(j}l§

Fig. 1.43. The EOCI-pool illustrated as an "Ice-berg". From Hakanson and Jonsson (1989).

In sediments. onJy about 2 % of a sum-parameter like TOCl consists of EOCl. and only a small fraction « 1%)
of EOCI consists of chemicaJly identified substances. specific parameters. like dioxins. PCBs or DDTs. This is
sometimes illustrated by means of an "ice-berg" analogy (fig. 1.43). The toxicity of the unidentified part of EOCI.
or other sum-parameters and complex waste waters. may. however. be evaJuated by different methods (fig. 1.43B)
like:

1. the propensity for bioaccumulauon - the bigber the bioaccumulation. the greater the potentiaJ toxicity
2. the lipofIlicity - the bigher. the greater the potentiaJ toxicity
3. tile affinity for carrier particles (like humic materiaJ. FelMn-oxides and hydroxides. seston) - the lower the
affinity. the greater the potentiaJ toxicity
4. the molecular weight - the smaller. the greater the potentiaJ toxicity
5. the persistence - the easily degradable substances have lower potentiaJ toxicity.

The toxicity may be manifested in many different ways. e.g .• increased fin erosion in perch. increased frequencies
of skin ulcers in herring and increased skeleton deformations, like deformed jaws in pike or spinaJ column
bending in fourhom sculpin (Bengtsson, 1991), (fig. 1.44). SOdergren (1992) gives a thorough evaJuation of
biological effects (not ecosystem effects) of bleached pulp mill effluents in the BaJtic.

1.5.2. Effect-load-sensitivity

Fig. 1,45 illustrates some important questions that must be addressed if ELS-models for organic toxins. in this
example EOCI, are to be developed. The Steps in the environmenl2.l consequence anaJysis start from tile chosen
effect variables, and what they represent. All load and sensitivity variables must be adjusted to the time and area
resolution vaJid for the effect variable. If. for example, one chooses perch as the biological indicator organism

62
Contamination
Chlorinated organics

EOCI
in surfldal sediments
IIom A~arc:as 1r~ Wg IG

'4f4=!!L0=.
-~~ ~<:
l:::::~----.--)~._.:.~-'iIlii";""'''4\)''''
.•.

0-.,
EOCI
1000 """"
2000
lBOO
'V 6000 o D1ox1nS (pgIg erg.)
1600 c"""",,, (pgIg"'l!')
5000 1400 • EOCl (vgfg erg.)
4000 1200
1000
5000
BOO
2000 6
400
1000
200

0 5 10 15 20 2S 30 55
J::)bo!,::ancetr-cm 199=md (lan)

Fig. 1.44. Areal distribution of EOCI in surficial A-sediments in the Baltic, illustration of various effects of
chlorinated organic in fish (fm erosion, skin ulcers and deformed skeleton), and a diagram (simplified from
Jonsson et al., 1993) illustrating the relationship between EOCI, dioxins and furanes in surficial A-sediments
taken at different distances from the Iggesund paper and pulp mill towards the open Bothnian Sea. Based on
information in Hiikanson et al. (l988b).

(fig. 1.45) and then, e.g., uses the EOCI-content in perch muscle, the liver somatic index (LSI) or physiological
parameters, such as cytochrome P 54 activity in liver as effect variables, it is important to known the time and
area resolution of these variables for perch.

It is also important to note that one must be able to determine all variables of interest in a scientifica1Jy relevant
and, preferably, simple manner. It is well-known that many of the interesting sensitivity variables, such as water
temperature, may vary considerably within a coastal area and with time. Then, one Single value has little
representativity and gives a low predictive power relative to an effect variable. Obtaining effect, load and

63
EOCI
Mesocosm tests

E voriables
Crlt. conc.
Effect variables linked to perch
.. Physiological key pClram., e.g.
cy~ochrome P S4
.. Biologie,al key p.aram., e.g. LSI
(1iver-som.tie index)
.. EOCl in muse Ie ?
Area resolution: Entire bay (area<
10ian2.
Time rtsolution; 3-5 months
ing season)

Other effect vwiables


.. $ defect Pontoporiea
*$ reduction of brown algu
Aru resolution: 1
Time resolution :1

Loadva.riables: Sensiti'rityvwiables:
A. EOCI in sediment trap m.terial .. 't(..~ir ch.miru"y (nliniiy, te-mp.)
(or :>Urfici.l sedimen~s) ,. Intern.l1o.ding/botiom dy,,,mics!
rtslJspension (BA, BT &
B. EPOel, dioxins, furanes, .. \t'.te-r ntfflllon (Stlrface w.tt'r &
nals, DDTs, PCBs & HCB in s.clim'n!' bottom ",.te-r)
trap material.

?
?
Fig. 1.45. Examples of possible effect, load and sensitivity variables for EOCI in marine areas. From Hiikanson
(1990).

sensitivity variables for sum-parameters, like EOCl, andlor specific parameters belonging to the group of
chlorinated or halogenated organic material (fig. 1,45), is evidently a very impertant and comprehensive research
IaSk.

It is probable that most of the substances making up the EOCl-peol are rather conservative in the water mass, that
they have a long degradation time and are distributed over vast areas. Considering the significant load of
chlorinated organic substances to the Gulf of Bothnia (about 11,000 tons TOCl/year in the middle of the 1980s,
fig. 1.46), the probable high sensitivity of the area to many of these substances (e.g., Ule theoretical water
turnover time in the Baltic is very long, about 35 years, which means that a pollutant could remain in the system
for a very long period of time), to a number of established biological effects on individual animals (see
SOdergren, 1992), to the size of the impact area (fig. 1.47) and the probable duration of the contamination, one
must conclude that there are many arguments for a position where the emissions from the paper and pulp mills
(PPMs) and other halogenated organics are considered as a major ecologicalthreaL

64
Husum
2040
DOIll:sjl:.i
no A.

tl:strand
960

19gesund
1080 Gulf or Bothni.
lJallvll<.
525
Norrsundet
no
Korsna:s - •
1790 Sl<.utskar
\JOO

Q lQO lOll

7000;---------------------------------------,
B.
6000 Icg(EOCl) - 2.57 + O.44'log(IOCl/dist:A Z)
for TOCI. 2040 t/yr for Husum
5000
o
;;, 4000

~ SOOO
o
"' 2000

1000 EOClcrlt. 500


If-
°O~--~~--~~~--~~~------------~
2.5 5 7.5 10 12.5 15 17.5 20 22.5 25
Dtstance (km)

70
y--l.65x+47.2;,.2 -0.52:n-16:p -0.0016 C.
0
"- 60 •
i!:.
• 50 •
]u • • •
"!l 40 • •

il 8
""
.E
SO

~ 25%
.c 20 •
••c 10



t.x 0
0 2.5 5 7.5 10 12.5 15 17.5 20 22.5 25
D1stance (km)

Fig. 1.46. A. Emissions of total organically bound chlorine (TOO) in tons/yr to the Bothnian Sea in the middle
of the 1980s. From Hilkanson et al. (1988b).
B. The relationship between EOCI-concentrations in surface sediments (0-1 cm) and distance from the Husum
PPM. which emilled 2040 tonslyr of TOO in the mid-1980s. The equation describing the relationship is from
Hilkanson et al. (l988b). The "critical" limit is estimated from the information in fig. C.
C. The relationship (regression line and statistics) between the percentage of fourhorn sculpins showing severe
defects and distance from the Husum PPM (using data from Bengtsson, 1991). The "critical" EOCI-
concentration (Eelit) of 500 Jlg EOClIg IG (IG = loss on ignition = organic content) is here related to a 25%
level in defect sculpins corresponding to a 2.5 fold increase as compared to the assumed reference level of 10%
defect sculpins.

6S
Fig. 1.47. Extractable organic chlorine (EOCI in l1g per g organic material) in surficial sediments (0-1 cm)
from the Baltic Sea, Kattegat and Skagerrak. TIle highest value is 6270. From Jonsson (1992).

One approach to estimate a "critical" limit when E is smaller than 10 and very lillie information is available
concerning causal relationships between load and effect variables is illustrated in fig. 1.46. The "critical" EOO-
level is here defined as 500 l1g EOCI per gram organic matter (loss on ignition) of surficial (0 - I ern) sediments
related to a 25% level in defect sculpins corresponding to a 2.5 fold increase as compared to an reference level of
10% defect sculpins. This definition of Ecrit is crude, based on a statistical analysis and not founded in
mechanistic knowledge Or in validated ELS-models. Lacking such information, this is a temporary operational
solution. The argument for this is that it is better to give such definitions and thereby clarify the limited
quantitative knowledge on the given subjec~ than to cover this by speculations about, e.g., causes and effects
based on analogies valid for other substances in other environments. This definition of Ecrit for the PER-analysis
is also objective in the sense that it can be used by all in the same way. This defmition implies a chalienge to
make improvements.

1.5.3, Geographical perspective

Fig. 1.44 illustrates the geographical distribution pattern of EOCI and specific toxins (dioxins and furanes) in
Baltic sediments. How much of the EOCI was found in the sediments? It has been estimated (Hiikanson et al.,
1988b) that about III 0 of the discharge of EOCI to the southern part of the Bothnian Sea could be found in the
sediments of this area; between 5 and 50 tons EOCI were, e.g., found in the coastal sediments outside Iggesund
(fig. 1.44). From fig. 1.44 one may also note:

(I) There is a large-scale spread of EOO from Swedish PPMs. The larger the emissions (see fig. 1.46), the larger
the impact areas in the Baltic (see map in fig. 1.44; fig. 1.47 gives a more detailed map of the areal distribution of
EOO in sediments surrounding Sweden), the greater the potential ecological problems.

(2) The net coastal current acts as a barrier to the spread of EOCI and related substances and the main direction of
transpon of the substances is not primarily towards the centre of Ille Boillnian Sea but in a predominantly souillem
direction along the Swedish coast (see fig. 1.48).

66
Area of hiOh
contamination

'- J
,
~C::,,/'
I
/.ffr \
1
1
~ Area of Intermediate
contamination
.~/
_ I Area ot \\
.::: I

:;:
0,
<.J 1
dlffu8e
Influence -
\

'\
\
I

( \ tow-Intermedlate)
~ " contamination I Area of low-Intermediate
'.... / ....,J.--- contamination

~ {·:·~·fI
Fig. 1.48. l11ustration of ille coastal jet-zone and ille major hydrological flow pattern in ille Boillnian Bay and
ille Boillnian Sea. From Hakanson et al. (1988b).

The following features should be, ~oted:


• The dominating water circulation in each basin (ille Boihnian Bay, ille Boillnian Sea and tile Baltic Proper)
constitutes an anti-clockwise cell, which distributes ille settling particles, ille suspended material and ille
poliutanlS in a typical pattern, reflecting ille flow of ille water. This anti-clockwise cell is created by ille rotation of
Ole earOl (Ole Coriolis force), which defleclS any plume of flowing water to ille right in relation to tile direction of
ille flow in Ole norillern hemisphere (and to ille left in ille souillern hemisphere). Thus, when a Swedish river
enters Ole Baltic, tile water turns to tile right and follows ille shore.
• The net hydrological flow is to tile SOUtil on ille west (Swedish) side of ille Baltic.
• The currenlS are railler strong and stable close to land and weaker towards ille center of each basin.
• The figure illustrates only ille net component of ille flow - this means illat ille water would also flow in most
oiller directions during tile year.

(3) The contamination is greater on tile Swedish side illan on tile Finnish side because of greater Swedish
emissions of chlorinated organic substances from PPMs (see fig. 1.47).

So, Ole distribution of EOCl in surficial sedimenlS (0 - 1 em) sbows a very characteristic pattern wiill high
concentrations close to Ole PPMs, low concentrations in areas of erosion and transport, out to about 10 - 15 km
from ille coast, and comparatively high concentrations again in open water areas wiillin ille zones of continuous
sediment accumulation.

67
o

r-'A"~i~"",,-~--------~~----~---oP'nS"-----------+------------

E· and !- '"U"nu,"

E- and !-'"U,m,"

A-bottom
( line p;utiele~ J

Fig. 1.49. Bottom dynamic conditions outside Norrsundet in the Bothnian Sea. TIle emissions from this paper
and pulp mill may be distributed in many cycles. The accumulation areas in the open water part of the
Bothnian Sea appear at water depth below 50 m. Sediment sample from shallower areas, often tJansportation
areas, may vary very much, and be very old. From Hakanson et al. (1988b).

68
Fig. 1.49 illustrates why it is importm1t to lmow the bottom dynamic conditions in contexts dealing with the
spread of pollutants from point sources (in this example from the paper and pulp mill at Norrsundet on the
Swedish coast in the Bothnian Sea). The fine materials (clays, humus, seston, etc.), which have a strong capacity
to bind various types of pollutants, like Chlorinated organic material (PCBs, dioxins, etc.), will be transported via
several resuspension cycles from the site of emission to the true accumulation area at water depths greater than the
wave base at about 50 m in the Bothnian Bay. The 19gesund area has, in contrast to the area outside Norrsundet in
fig. 1.49, well-defmed accumulation areas, and the diagram in fig. 1.44 shows that the concentrations of EOCl,
dioxins and furanes increase markedly towards the industry, which is the only lmown anthropogenic point source
of chlorinated substances in the area The sediment concentrations of certain dioxins, furanes and guaiacoles
correlate markedly with EOCI. The chemical, structural characteristics of the dioxins and furanes correspond to
the pattern lmown to be specific for certain PPMs.

Fig. 1.50 gives an over-all mass-balance for EOCI for the entire Baltic for 1988-89. It is based on direct
measurements and emission data. The emissions from PPMs clearly dominated the input (75%), while the inflow
from tributary rivers was very small « 2%). The atmospheric input was about 23% of the total inpuL More than
50% of the total input since the early 1940s was stored in the Baltic system. So, the retention time of the
substances in the Baltic system is long, and this brings us to the next PER-criteria the time perspective.

1.5.4. Time perspective

A large number of sediment cores have been taken from Baltic accumulation areas and analysed for EOCI. Fig.
1.51 (upper) gives results reflecting the average condition. The increase in the EOCI concentration in the
sediments, and hence the increase in sediment and water contamination, started in the late 1950s and coincide well
with the general contamination and increase in eutrophication in the Baltic. In the next section, we will discuss the
eutrophication problem.

It should be noted that most of the emissions ofTOCI to the Baltic (see fig. 1.46) have now been halted as a
response of legislation and public awareness of the problem (based on results published, e.g., by SOdergren et
al., 1988). The Baltic is slowly recovering from this contamination. The time perspective of this recovery could be
illustrated using data on two well-lmown organic pollutants, DDE and PCB (fig. 1.51, lower). There has been a
ban of these pollutants and the conditions in the Baltic are now improving. The ecological halflife for DDE and
PCB in herring in the Baltic is about 6 years. One can expect that such ecological halflives would also be valid for
many organic toxins emitted from PPMs. This means that it will probably take 20-40 years until the system has
recovered so that the operational E-vanable, tile EOCI-concentration in surficial sediments. is back to Ecrit = 500
~g EOCUg org. matter. The PCB-levels in aquatic biota are related to the net budgets to the entire ecosystem, and
there are still many questions in that bud geL notably concerning the organic pollutants stored in the sediments
(Fig. 1.51, lower). A significant part of the small sized, volatile toxins will probably leave the Baltic ecosystem
via the water surface to the atmosphere, while a larger part of the large-molecular compounds will be deposited in
the sediments. But what will happen to the pollutants stored in the sediments if the environmental conditions
change, e.g., if the load of nutrients is reduced? This question is now being addressed by a major Swedish
research project (the EUCON projecL see Jonsson, 1996).

69
Atmosphere Atmosphere Atmosphere
Pulp mills. Pulp mills. Pulp mills.
Rivers I 105 Rivers I 28 Rivers I 14
• 100 •1 350 <150 •
... 1
Kottegott Water
_1
Sediment

Baltic Bothnian Bothnian


propElr Sea Bay
Fig. 1.50. A budget for EOCI for the three sub-basins of the Baltic Sea based on data for 1988-1989. The
fluxes (arrows) are given in tonslyr and the amounts in tons. From Jonsson (1992).

Veer
a ;;:::;r=t 1965

1970

' - - - - - -----+(1950)

Q,) -10
>
'"
-"

-20 +---_---,-___---1
a 10 20
EOCI (ug/g ds)

DOE PCB
1.0 -
0.9 o
o
0.8
o
0.7 0

0.6
0,5 0

o o
0,4 o
o
0,3
0 0
0,2 o

0,1

79 81 83 85 87 89 91 93

Fig. 1.51. Upper. The average historical development for EOCI in the Baltic Proper (based on 8 sediment cores
and 38 surficial sediment samples). From Jonsson (1992).
Lower. The concentration of DDE and PCB in herring muscle (values in Ilg/g fat) from the Landson station in
the Baltic Proper for the years 1979 to 1993. From Ostersjo (1993).

70
1.5.5. Summary - chlorinated organics

• No ELS-models are available for chlorinated organics for aquatic ecosystems. There are strong indications of
dnmage to the reproduction of seais (from planar PCBs). and indications of damage to individual fish (spinal
column bending. deformed jaws. skin ulcers. etc.). but only indications of dnmage to the strUcture of the
ecosystems (see SOdergren. 1992). Since little is known about causal linkages between load. sensitivity and effect
variables for coastal ecosystems. it is very difficult to set a relevant E-value in this instance. However. the
potential ecosystems effects are great, and by applying the precautionary principle. the E-value is set to 6 in tilis
approach ("clear real ecosystems effects"). This value could. and should. be challenged.
• The contamination has a wide geographical spread in the Baltic (Fig. 1.47). The A-value is set to 6 - coastal
areas in many regions with E higher than the Ecrit-limit of 500!lg EOCI per g org. material in surface sediments.
• The contamination will likely continue for several decades. The T-value is set to 6 ("20 - 40 years").
• It is difficult to evaluate PER for this contariiination. In this approach. it is 216. but there are great uncertainties
about the E-value and no validated ELS-models. Thus. the PER-analysis can be used to motivate why this ought
to be a topic for high priority research.

1.6. Nutrients

1.6.1. Introduction to eutrophication

Richard Vollenweider (Vollenweider. 1968. 1976) presented the first load models for phosphorus for lakes in the
late 1960s. Water management was then at a rudimentary stage dominated by descriptions. verbal "models". and
elaborate "logical explanations". The practical usefulness of most results and models were often negligible. as was
the predictive power. Aquatic ecosystems were rightly recognised as extremely complex. and this gave legitimacy
to the predictive failure. The argument was that very many .factors could influence the primary production of a lake.
many different nutrients and chemical forms of nutrients ought to be considered. differently in different lakes.
different seasons. for different species of algae and plankton. and so on. All this had to be studied before anything
useful could be said or predicted about the status of the ecosystem. Different ways to halt and reverse
eutrophication were suggested. but many treatments were expensive. speculative and untested. Vollenweider
approached the problem the other way around - he tried to simplify! By means of simple mass-balance calculations
and statistical regressions (see section 2) using what then seemed a large sel of empirical data. he could
demonstrate that in many lakes. eutrophication could be reversed by reducing the input of total phosphorus(TP) to
the lakes so that the mean lake annual concentration ofTP(notfraction so-or-so) could be lowered. At flfst, this
was met with scepticism. But his results proved to have predictive power and practical applicability. Since then.
many studies have demonstrated where the Vollenweider approach can - and cannot - be used. Different alternative
models have been presented. and the most successful of those have one thing in common with the basic
Vollenweider model- simplicity!

Thus. total phosphorus is since long recognised as the most crucial limiting nutrient for lake primary production
(Ahlgren. 1970; Schindler. 1977, 1978; Bierman. 1980; Chapra. 1980; Boynton et al.. 1982; Wetzel. 1983; Persson
and Jansson. 1988; Boers et al .. 1993). The literature on phosphorus in lakes is extensive. Nitrogen is generally
recognised at the most limiting nutrient in marine areas. and both elements are of vital importance in estuaries and
brackish waters. like the Baltic (Redfield. 1958; Ryther and Dunstan. 1971; Myers and Iverson. 1981; Nixon and
Pilson. 1983; Howarth and Cole. 1985; Howarth. 1988; Hecky and Kilharn. 1988; Ambia. 1990; Nixon. 1990).

71
23.8 19.2 12.4 7.5
1000
Tot-? lake 6.25
(mg/nill 5.1
3.8
100
2.6
%: 80
SO
2.1
"t;;,100 1.4
g 30
25
~ 20 /0.7
0
15
E
0.
10
8
Chl-a
(mg/m3 )
, 10
~
5-----
{:.
2.5

0.1 1 10 100 1000


Water residence time, Tw (years)

Fig. 1.52. The ~EeD-model for lake eutrophication illustrated as a load diagram. TIle diagram relates tributary
inflow concentrations of total phosphorus, to theoretical lake water retention time, lake concentrations of total
phosphorus and lake concentration of chlorophyll. From Wallin et al. (1992) based on OEeD (1982).

The famous Vollenweider model (Vollenweider, 1968, 1976; and later versions, e.g., OEeD, 1982, see figures 1.52
and 1.64), and the analysis behind this load model, constitutes a fundamental base for practically all environmental
assessments of phosphorus in lakes. This approach has played a paramount role in lake management, but mese
concepts have not yet fully penetrated into marine or terrestrial ecology. The main reason for this may be mat
limnologists are trained to have an ecosystem perspective, i.e., to study entire lakes. In lake models, me effect
variables are generally linked to the concepts of hypertrophy (extremely productive lakes), eutrophy,
mesotrophy and oligotrophy, which are related to defined mean annual concentrations of total phosphorus in me
lake water (table 1.5A).

The interesting pan is not to predict a simple mean concentration of a chentical element like total phosphorus, but
to predict ecological effects related to lake TP. From table 1.6, however, it is clear that lake TP can be
quantitatively related to almost any ecological effect variable or key functional group characterising lake
ecosystems. Fig. 1.53 gives a list of such ecological effect variables related to nutrients: Secchi depth, chlorophyll-
a, bottom fauna state index, fish state index, hypolimnetic oxygen demand and oxygen concentration, etc. It is
evident that the concentration of lake TP can be influenced by emissions from many types of sources, like point
sources (e.g., domestic sewage, industries and fish farms), atmospheric deposition (to me lake surface and me
catchment area), intemal loading (linked to resuspension, diffusion, etc.)and,oftenmostimportant, tributary input.
The characteristics of the catchment, like bedrocks, soils, land use, etc., regulate the concentration of TP in me
tributaries to tile lake (see section 7, which presents the LEEDS model for lake eutrophication).

72
Table 1.5A. Characteristic features in lakes of differentlrophic levels. Note that there is a great overlap
between tJle different categories, such that in oligolrophic lakes the concenlrations of LOtaI-P may vary within a
year from very low to high values (modified from Hiikanson and Jansson, 1983).
B. Trophic categories in Baltic coastal areas. All variables are expressed as summer mean values (from
Wallin et al., 1992). Abbreviations are given in table 1.8. From Hakanson (1994).

A.
Trophic Primary Secchi Chl-a Algal Total-P Total-N Dominant
level prod. (m) vol.*) (**) (**) fISh
(g C/m2 'yr) (mg/m 3) (glm 3 ) 3
---(mglm )----

OligOL <30 >5 <2.5 <0.8 <10 <350 Trout, Whitefish


MesoL 25-60 3-6 2-8 0.5-1.9 8-25 300-500 Whitefish, Perch
EUl. 40-200 1-4 . 6-35 1.2-2.5 20-100 350-600 Perch, Roach
Hypen. 130-600 0-2 30-400 2.1-20 >80 >600 Roach, Bream

• = Mean value for tJle growing period (May - Oct.)


** - Mean value for the spring circulation
B.
Trophic Secchi Chl-a Total-N Inorg-N SedS Ozll 02 Sat
level (m) 3
(mg/m ) 3
(mg/m ) (mg/m 3) 2
(glm 'd) (mgll) (%)

OligOL >6 <1 <260 <10 <2 >10 >90


Mesal. 3-6 1-3 260-350 10-30 2-10 6-10 60-90
EUl. 1.5-3 3-5 350-400 30-40 10-15 4-6 40-60
Hypert. <1.5 >5 AOO AO >15 <4 <40

1.6.2. Marine eutrophication

There are positive as well as negative effects of aquatic eutrophication. Positive effects are, e.g., increases catches
of fish in the Baltic (fig. 1.54). Note, however, that tJle catch per unit effon has decreased since the late 1970s (fig.
1.55). Negative effects are, e.g., decreased caches for certain species of fish and altered fish communities (see, e.g.,
Kautsky and Kautsky, 1989; Hngren, 1989; Hansson and Rudstam, 1990; KautsJ..-y, 1991), changes in the structure
and composition of otller key, functional groups of organisms, like large perennial algae (FlIcus vesiclI/oslIs). It
should be noted tJlat tlle important cod population in the Baltic bas decreased significantly since the early 1980's.
Why has tllis happened? It cena.inly depends on many factors in a complicated system of interactions, but one of
those is linked to the oxygen conditions (Voipio, 198 I).

Each female cod lays many eggs. In salty water, the roe remains suspended in Ibe surface, oxygen-rich, water. In
the brackish surface waters of tlle Baltic, tlle salinity and hence also the density is not high enough to allow the roe
to floaL It sinks to the saline bottom water before Ibis descent is stopped by Ibe greater density of Ibe water. If Ibis
bottom water is oxygen-rich, tllere will be a good spawning year for cod. but if tlle oxygen content is low, Ibe eggs
will die and the spawning will fail.

73
Target effect Ta..tiablet

Operational
effect indicators

Fig. 1.53. The character of the drainage area influences lake water qUality. This is a statement that is simple to
make, but how could it be quantified? It is evident that the geology, hydrology, land use and precipitation
influence the inputs of substances to lakes, including the key nument in lakes, phosphorus. It is an important
task to relate drainage area characteristics to lake characteristics and to predict variables of primary biological
importance, like lake total-P, which in tum may be related to operational ecological effect variables (like
oxygen concentration), and to the target effect variables (like reproduction and abundance of key species).
From Hiikanson (1994).

Not since 1977 have events of salt water inflows managed to improve the oxygen siruation in the deep areas in the
Baltic Proper for any longer periods. The production of organic material, algae, plankton, bacteria, animals, has
been so great that the addition of oxygen that still enters from the Kattegat is quickly utilized in the degradation of
this continuously increasing "biological rain" down in the deep areas.

The surface water is generally well oxygenated with 02-concentrations between 7 and 9 mill; the values from
January are often higher than for September depending on the bigh production of organic material during the
summer months. When the algae and plankton die, they are consumed by bacteria and in this
decomposition/mineralization of dead organic matter oxygen is utilized (fig. \.56). This results in lower 02-valueS
in September. From about 60-70 m water depth, one can generally note a very steep 02-gradient; and at a water
depth of about 140 m, there is often no oxygen at all left; this is the so-called redoxcline. Beneath the redoxcline,
hydrogen sulphide (I-!2S; a compound that smells like rotten eggs) is produced in microbiological and chemical
processes. Fig. 1.57 illustrates that the areas with laminated sediments, where the oxygen concentration is so low «
2 mgll) that the bottom fauna die, bave increases very much during the eutrophication processes in the Baltic, from
1960 and onwards.

74
Table 1.6. Selected regressions illustrating ibe key role of lake tolal phosphorus in Predictive limnology. Many
biological variables that would nonnally require detenninations from extensive and expensive field and
laboratory work may be eslimatedlpredicted from one key, abiotic state variable, tolal-P (TP in mg/m 3). Some
variables may be predicted with great accuracy, oibers with low. N = number of lakes used in the regression.
ww = wet weight. dw = dry weighL From Peters (1986) and Hitkanson and Peters (1995).

y-value Eguation Range forTP r2.value N Units


r20.0.7
Chlorophyll y=o.073*TP1.44 3-300 0.9 77 mg/m 3
Max. prim. prod. y=20*TP-71 7-200 0.95 38 mgClm 3 *d
Mean prim. prod. y=1O*TP-79 8-200 0.94 38 mgC/m 3 *d
Nanoplaakton y=17*TP1.3 3-80 0.93 23 mgww/m3
Phytoplankton y=30*TPIA 3-80 0.88 27 mgww/m3
Fish yield y':'7.1 *TP " "iO-550 0.87 21 mg wwfm2 *yr
Macrozooplaakton y=20*TPO.65 3-80 0.86 12 mgww/m3
Zooplaakton y=38*TPO.64 3-80 0.86 12 mgww/m3
Bacteria y=O.90*TPO.66 3-60 0.83 12 milLimi
Net plankton y=8.6*TPL7 3-80 0.82 23 mgww/m3
Fish y=590*~·71 10-550 0.75 18 mgww/m 2
Crustacean plankton y=5.7*TpO.91 3-300 0.72 49 mgdw/m3
Microzooplankton y=17*TPO.71 3-80 0.72 12 mgww/m3
Blue-greens y=43*TPO.98 8-1300 0.71 29 mgww/m3
r20.0.25
Benthos y=810*~·71 3-100 OA8 38 mgww/m 2
r 2<0.25
Transparency y=O.8*TP-0.28 3-400 0.22 87 m

Numerous reports during the early 1950s demonstrate that the higher animal life in the bottoms in large areas was
being eliminated as a result of oxygen deficiency. In those parts laminated sediments were encountered (Jonsson,
1992). Laminated deposits dominate in anaerobic sediments, Le., in oxygen-depleted sediments, where mixing due
to bioturbation is negligible due to the extinction of the bottom fauna. TIle formation of anaerobic sediments is a
natural process occurring predominantly in deep, StagnanL stratified and highly productive aquatic systems. But
this natural phenomenon is now happening more often and in larger areas due to the anthropogenic eutrophication
of the Baltic system. Fine, darkish, thin layers are generally laid down during winter, and thicker browner layers
during summer. By counting the number of annual layers in sediment cores from the Baltic, it has been shown how
the oxygen-free bottoms have grown during recent decades. It started during ibe 1940s, but the particularly large
changes did not occur until the 1960-1970s (fig. \.58). The sediments also provide evidence of the increasingly
difficult conditions in which the bottom animals live. In the sediments from ibis period, it is often possible to record
the transition period between homogeneous and oxygenated sediments and those which are varvedllaminated and
oxygen-free. The bottoms are now periodically covered by extensive carpets of sulpburous bacteria. Sediment
removed from such places generally has an unpleasant odour of hydrogen sulphide, a gas which is toxic to higher
life.

75
'"z TOTAl.
---HE1UUNG
~ .........••.•• COD
------- SPRAT

...
~
Z
Z
-< - ..
~
1910 -20 -30 -40 -50 -60 -70 -80
Fig. 1.54. Catches of herring. sprat and cod in the Baltic Sea. Modified from Hansson and Rudstam (1990).
1S0

/0
10(0 .,....,

120

lOa

20

1979 1980 1981 1982 1983 198.( 19G~ 1986 19S7 19a8

Year
Fig. 1.55. Catches per unit effort (kg/hour) estimated as the mean from September to December. for cod taken
by cod bottom trawl (0) and lobster Irawl (.) in SE Kattegat from 1978 to 1988. Modified from Baden et aI.
(1990).

Oxic sediments generally have an abundant macrofauna which cause a mixing of the sediments. The result is a
mare or less homogeneous mixture with a colour that varies in different shades of brown (from grey/hrown to
blackihrown). When oxygen·rich water returns to the bottoms. the bottom fauna can s!art to recolonize the
sediments. The laminated deposits may then be more or less homogenised, and so on in cycle after cycle. Fig. 1.58
gives a compilation of several changes in the Ballic related to eUlrophicalion, increased catch of fish, decreased
oxygen conditions in the bottom water, increased areas with no bioturbation ("dead bottoms") and increased
sedimentation of organic materials and nutrients.

76
of organic matter
Surfllce wtIter

Deep wnter
Increased oxygen
consumption

Increased sedi-
mentation of
contaminants
associated to
I
Decreased oxygen
concentration
I
Decreased benthic
"" Occurence
of H~
I
Precipitation
macrofauna of sulphide
"","'om~ I binding mer.",s,
Decreased minera-
lization rate

Fig. 1.56. illustration of the sequence from increased primary production (eutrophication) to increased
sedimentation an increase entrapment (= sequestering) of contaminants in Baltic sediments. From Jonsson
(1992).

Area (k"12)
80000.,----------------,
iiii Recently Laminated
iiiii sediments
60000 § Naturally laminated
E::j sediments

40000

20000

1900 1920 1940 1960 1980

Fig. 1.57. Schematic illustration of the probable expansion of laminated sediments in the Baltic Proper duling
the last century. From Jonsson (1992).

77
o Total fish catch
db Oxygen at 300 m, Landsort Deep
- Area of laminated sediments
A N03-N, surface water, Gotland Deep

® Organic content in surficial sediment


lilliI Sediment sequestering of carbon
deriving from net production

€l

• •
I "" I • j •

1900 1910 1920 1930 1940 1950 1960 1970 1980 1990
, • Iii ·t i.. I

Fig. 1.58. Recent eutrophication record for the Baltic Proper on a relative scale as exemplified by the total fish
catch (max. value = 900,000 tons), oxygen concentration (max. value 2.8 mUl), area of laminated sediments
(max. area 70,000 km 2), nitrate concentration in surface water (max. value 10 J.UIlollI), organic content of
surficial sediments (max. value 16.8 %) and sediment sequestering of carbon deriving from net bioproduction
(max. value 9 g C/m 2'yr). From Jonsson (1992).

Table 1.7. Transport in 1985, 1987 and 1990 of nutrients from Sweden to coastal areas and related
environmental goals. From SNV (1993b).

Total nitrogen.(tons/yr) Total phosphorus (tonslyr)


1985 1987 1990 1985 1987 1990
From Sweden

Rivers, natural 50600 50600 50600 2060 2060 2060


Rivers, anthropogenous 62700 54000 47370 1400 1460 1360
Urban coastal emissions 14620 14300 15150 835 770 575
Coastal industrial emissions 3170 4500 4315 850 750 545
Coastal fish fanus 180 200 405 25 30 60

Iinput 131270 123600 117840 5170 5070 4600

Swedish goal 101200 4120


HELCOMgoal 87100 3565

78
The major Swedish sources for Ule marine eutrophication in the Baltic are given in table 1.7. Among the given
sources. Ule input from rivers is the dominating one. However. for the Baltic there are also other sources and the
atmospheric nitrification (= nitrogen fIXation) is the single most important flux of nitrogen to the Baltic. and the
nutrient transpon from other countries. like Poland. is much greater than the Swedish contribution. Today Sweden
is far from the goal set by HELCOM (the Helsinki commission) for the load of nutrients to the Baltic.

1.6.2.1. Effect-load-sensitivity

The first ELS-models for nutrients for marine coastal ecosystems were presented by Wallin and Hiikanson (1991).
Examples of standard effect, load and sensitivity variables for the Baltic are given in table 1.8. The data used in the
development of the following coastal models in this book are all related to the information in this table.

An ELS-model (using chlorophyll as effect variable) is illustrated in fig. 1.59.lt should be stressed that it is NOT
possible to make simple adjustments of the Vollenweider model for lakes to marine areas. because nitrogen must
also be considered (Rosenberg. 1985; Ambio. 1990) and because the physical conditions of the sea often have a
profound influence on coastal ecosystems. A typical water retention time for lakes is about I year; for coastal areas
it is about 2-4 days! A low mean depth in a lake implies significant resuspension and a high internal loading; in an
open coast, it implies a low internal loading of nutrients since the bottoms would be dominated by coarse and hard
deposits WiUI a low content of nutrients.

For the model in fig. 1.59. one can note that the effect variable (chlorophyll) increases when the N-concentration
increases and that the r 2-value between these two variables is very bigh. 0.91 (based on data for 23 Baltic coastal
areas for the production period when the highest values of this effect variable are most likely to appear). Also note
that the total-N concentration is called a response variable because it reflects the total ecosystem response of all the
inflow (load) and outflow fluxes of nitrogen for a given coastal area. as well as all types of internal processes. It is
much easier to measure the TN-concentration Ulan it is to measure and model all the processes regulating the given
TN-concentration. and Ule TN-concentration is a typical "collective" variable in this context of predictive ecology -
it is relative easy to measure. it gives a very good prediction of a target effect varinble. it is mechanistically sound.
it reflects very many complicated processes. and it is a standard variable in coastal monitoring and research.

Anotiler ELS-model for coastal eutrophication is presented in fig. 1.60. The mean summer oxygen saturation in the
bottom water (i.e .• Ule water beneath the thermocline) is used as a simple operational effect variable (02Sat). The
rationale for using this chemical measure is given in fig. 1.61. which illustrates the relationship between bottom
fauna. oxygen conditions and load of organic matter to the sediments. When the 02-saturation is lower than about
20%. key functional benthic groups are extinct. It is. naturally. convenient to develop ELS-models based on such
simple chemical effect variables as tile 02Sat, but then one must demonstrate the biological/ecological Significance
of such effect variables. This is quite clear for 02Sat (fig. 1.61).

The following working hypotheses were tested and verified when Ule model in fig. 1.60 was constructed (all steps
of the model construction are given in section 2):

1. No single factor could explain the variability in mean 02Sat that exists among these 23 coastal areas. Several
factors must be included and each factor is likely to add only a limited predictive power.

79
, 0
ELS-diagram - ti:a
Nitrogen in Baltic coastal areas !':'"
::g
5 1.5 15 3 30 450 50
5,-----------------------~-----
4,5 0

4
:a
~g
:2'" 3.5 3 3 10 6 60 350 30
U 3
b" 2,5

~ 2
tll 1,5
1 6 2 10 90 250 10
0,5
O+-rT-r~rT_r~rT_rTO_r._TO_i ~ ~
.5 .5
z z
-;it
'0
l-
"
.~

I:."
0
.5
Mode! (see sectiOD 2.4.2):

!og(Chl) = 2.78'!og(TN) - 6.66

u
= =
(1"2 0.91, n 22)

Legeod and presuppositions:

Response variable: TN :::::total nitrogen concent:ration in mg/m3;


range: 250 to 420;
t Effect variables
t Load variables
TN is a function of the total load of nitrogen to the coastal area,
and of the internal processes
Model not applicable for:
Effect variable: Mean chlorophyll- a concentraIion in mg/mA3;
presuppositions: .. Lakes. bracldsh W31ers and esruari.es (where
• area.: 1 to 14 kro A2 also the load of phosphorus has to be accounted
.. time: the production period (Jv1ay to SepL) for)
.. Marine areas influenced by tides (dH>25 em)
.. Coastal areas outside the range afthe model
parameterS

Fig. 1.59. An ELS-model for coastal eutrophication where tbe concentration of chlorophyll-a is used as an
operational effect variable and tbe concentration of total-N (TN) as a load or response variable. The model is
also discussed in section 2.

2. 02Sat depends on botb load factors (table 1.8 gives a list of all such factors) and sensitivity factors linked to the
size and form (the morphometry) of tbe coast. "The way tbe coast looks regulates how the coast functions".

It should be noted that the mean 02Sat-value is NOT a constant. but a variable, and that a model based only on
morphomeuic parameters can NOT be used for site-specific predictions of time-dependent y-variables. The
empirical model presented in fig. 1.60 is based on nuuient concentrations (total-N and total-P; load or rather
response variables) and morphometry to predict long-time summer averages of 02Sat for entire and well-defined
coastal areas.

One very important question concerns the definition of the coastal ecosystem, i.e., where to place the
boundaries toward the sea andlor adjacent coastal areas. It is crucial to use a technique thal provides an ecologically
meaningful and practically useful defmilion of the coastal ecosys!em. The coastal boundaries are, in tilis context,
operationally defined by means of tbe "topographical bottle-neck !echnique" calculated from the exposure (or the
openness toward the sea and adjacent coastal areas), see fig. 1.6.

80
Table 1.8. Variables for eutrophication effect and load (concentrations are expressed as mean values for June-
September and load variables as mean values for 2 years), morphometric parameters and Statistics for 23 Baltic
coastal areas. From Hiikanson (1994).

Symbol Variable Units Mean Min. Max.


Eutrophication effects
Secc Secchi depth m 3.6 1.0 6.9
Chi near-surface chlorophyll-a mg om-3 2.76 0.90 9.60
SedS near-surface sedimentation gom-2o day-1 7.53 1.07 17.56
SedB near-boltom sedimentation gom-2odal" 1 25.85 5.31 82.53
02B near~boltom oxygen conc. mgol- 1 7.06 1.05 10.00
02Sat near-bottom 02 saturation % 67.0 8.5 98.1
Nutrient concentration
TN near-surface total-N m0o -m-3 335 256 417
IN near-surface inorganic-N m0oe m- 3 27 14 52
TP near-surface total-P m0oe m- 3 23 14 31
IP near-surface inorganic-P mg om- 3 4 2 12
Nutrient load
Ntot total N load kg Noy-1 14,979 3656 81,501
ANtot total N load (area-weighted) kg N okm- 2oy-1 4023 1532 19,585
PtOt total P load kg poy-1 1472 93 6956
APtOt total-P load (area-weighted) kg N okm- 2o y-1 492 27 3177
Size parameters
Dmax maximum depUl m 22.6 11.1 46.9
a water surface area km 2 4.71 1.05 14.15
Ab bottom area km 2 4.49 0.92 13.90
At section area towards the sea km 2 0.014 0.001 0.082
V water volume km 3 0.033 0.006 0.18
Form parameters
Dm mean depUl = Via m 7.6 3.8 13.8
xm mean slope % 4.83 2.21 8.17
Dr relative depUl = Dmax o,Jrr/20 o,Ja % 1.12 0.46 2.68
F shore irregularity % 190.5 104.2 507.4
Yd form factOr = 30DmlDmax 1.05 0.57 1.47
Special parameters
E exposure = 1000AtiAb 0.39 0.045 1.27
Ff filter factor km 3 6.73 0.059 30.71
MFf mean filter factor km 3 1.32 0.012 6.49
BA proportion of A-areas % 19.5 0 80.9
BET QroQortion of ET-areas % 80.5 19.1 100

81
For mean depth = 7.6 m. filter factor = 6.7 km3 and volume = 0.33 km3
&t 100
"0
0
1: 90
n>
0-
-ci 80 SENSITIVITY VARIABLE
0
'-
0- Theor. deep water ret. time

'"CO
..J

<t 1:
-"
n>
.c 70
C>

60 ,
(Td In deys)

'"><t '"'-'" '." . '.


'. '-
n> 50 '. '. , ,
r- ". ,
u
'"
u..
u..
"3:
0- 40
'.
'.
'.
'- '.
, ,
'"
n>
n>
'" 30
". '. "'16
'.

E '. '. '-


"
0
'.
12B 32 ... '. '- Crlttcal
::; 20 '. '-

-""'",
'-

~
10
02-setura\ion

'"0 0
o 0 0 0 0 0 0 0 a 0 0 0 0 0 0 0 0 000 0 0 000 0 0
~ ~ ~ ~ ~ ~ 0 _ N ~ ~ ~ ~ ~ ~ ~ 0 N ~ V ~ ~ ~ ~ ~ 0
v v 'It V 'q' 'q' In Vl til If) 1J) 1J) 1.1) 1J) If) tn ..a \0 >D \0 \0 \0 \0 \0 \0 ~ ~

RESPONSE FUNCTION
(TN + 100TP) (mg/m3)

r2-value
Step 1: Nutrient response function (TN+10"TP) 0.49
Step 2: Theor. deep water retention time (Td) 0.76
Step 3: Mean depth (Om) 0.89
Step 4: Fitter factor (Ff) 0.91
Step 5: Coastal volume (V) 0.93

Data from 23 Baltic coastal areas

Model:
02Sat= 1OO~SIN{ 14.827 -4 ,4648.~LOG(TN+ 1 O~T P)-O.403 "LOG{ 1 + Td)-1 ,04S"LOG(Dm)-O,021 ~FI+O,275·LOG(V))

Fig. 1.60. An ELS-model using the oxygen saturation in deep water, 02Sat, as opemtional effect variable for
Baltic coastal areas; critical concentration, "ladder" and model (see section 2 for model derivation).

The results shown in fig. 1.60 give 02Sat versus a load function plus sensitivity parameters. One can note that
02Sat may be predicted quite well from the load function (1N+ I O*TP) plus four morphometric (sensitivity)
parameters: The higher tile load of both N and p. the lower 02Sat; the longer the theoretical deep water retention
time (Td), the more sheltered the coast is (given by the filter factor, Ff; see fig. 1.63 for definition), the deeper and
larger tile coastal area and the lower the filter factor, the lower 02B. The r2-value after five steps is 0.93. The most
powerful predictor is the nutrient function.

Note that in models of this kind one can often replace the model variables (tlle x-variables) with related variables
from the same cluster or functional group, e.g., Ff may be replaced by tile section area; the load function
(1N+lO*TP) could be replaced by other types of load functions. Since "everything is related to everything else" in
ecosystems like these, it is generally difficult or impossible to give clear-cut causal explanations why a certain x-
variable is linked to a certain y-variable. Models should be built upon simple, easily accessible, logical variables
which show a minimum of inter-dependence.

82
Manne eutrophication
Nutrients
(phosphorus and r::-::71 Areas with low oxygen
nitrogen) ~ concentrations « 2 mgJU

~
Areas with lam1nated sedIments
and no bottom fauna

Finland
Sweden

3
(em)

Increased ccntamlnation of organic materials


Decreased oxygen concentration
Polen
Germany
Fig. 1.61. Illustration of the areal distribution of the eutrophication problem in marine areas surrounding
Sweden (based on Ambia, 1990); on the west coast, low 02-concentrations occasionally appear in bottom
water; on the east coast, laminated sediments occur over vast areas; and an illustration of ecological effeclS on
bottom fauna from increased organic load (oxygen consumption) in sedimenlS (based on Pearson and
Rosenberg, 1976). From Hakanson (1994).

The model in fig. 1.60 should NOT be used for other types of coaslS. It cannot, e.g., be used for coastal areas
dominated by heavy tides.

Many factors could potentially influence 02Sat, the effect variable in this case. It is easy 10 speculate and
qualitatively discuss such relationships. With empirical data it is possible to quantitatively rank such factors and
derive predictive models based on just a few, but the most important, faclOrs influencing 02Sat (for coasts of !be
given type). The given model could (statistically) explain 93% of the variability in the given y-variable among !be
23 coastal areas.

A "naturai" 02Sat could be estimated from the model, if it is possible to estimate "natural" (or reference)
background values of TN and TP. If the actual 02Sat-value of the coast differ from such a "natural" value, then
!bose divergences may be discussed in a quantitative manner.

83
Fig. 1.62. Extension of laminated surficial sediments in the Baltic Proper. Modified from Jonsson (1992).

1.6.2.2. The geographical perspective

Fig. 1.61 illustrates why marine eutropbication is sucb a problem: Very large areas along the Swedish west coast
bave bottom areas where the concentration of oxygen occasionally decreases below the critical limit (Eerit) of 2
mg/l, or 02Sat = 20%. A less scbematic map based on empirical measuremeDiS is given in fig. 1.62. Many benthic
animals die if the 02-concentration is lower than 2 mg/l (fig. 1.61 rigbt), bioturbation is halted, and laminated
sediments appear. The figure also sbows that very large areas in the Baltic Proper bave laminated sediments.
Recent studies (Per Jonsson, pers. comm.) bave also demonstrated that laminated sediments appear over larger and
larger areas also within the coastal zone. Tbis is alarming since the coastal zone is regarded (see Hakanson and
Rosenberg, 1985) as a "nursery and pantry" for the open water areas.

Tbe eutropbication conditions in the open water areas of the Baltic are governed by the tributary nutrient fluxes to
the Baltic, and by the bydrodynamic conditions, such as the coastal currents and the water exchange. The
conditions in the coastal zone are dynamic and strongly influenced by the conditions in the open water areas (a
typical water tlllUover time for a coastal area is 2-4 days). Also sbeltered bays deep within archipelago areas are
strongly influence by the bydro<lynamical conditions in the open water areas (Markus Meili, pcrs. comm.). This
means that one can only improve the conditions in the coastal areas by reducing the main nutrient fluxes from the
largest and most polluted rivers entering the Baltic.

84
Sea
central radial

Land

Definition of filter factor:


Ff= ICL(COS(ru)*X1)*At)J

Fig. 1.63. Illustration and definition of the filter factor (Fl). For further information, see Pilesj6 et aI. (1991).

1.6.2.3. The time scale

It is evidently very difficult to make a forecast about the future development of eutrophication in the Baltic. Tbe
theoretical water retention time of the Baltic is about 35 years (Voipio, 1981). The duration of the eutropbication
problem is not, bowever, primarily regulated by processes like water turnover time, sedimentation. biouptake and
internal loading but by the primary inputofnulrients via the main rivers. Tbese emissions are governed by land use
activities and discbarges from induslries and urban areas, and bence also by political decisions in the Baltic
counlries. In all likelihood, it will take 50 to 100 years to reauce the anthropogenic nulrient load to the Baltic since
sucb reductions bave implications{or agriculture, industry and society.

1.6.3. Summary - coastal eutrophication

• Coastal eutrophication (like lake eutrophication) can totally alter and eliminate key functional organisms.
Therefore, the E-value should be set to the maximum of 10.
• Coastal eutrophication impacts a very wide geographical area. The entire Baltic coast south of the Aland Sea is
clearly influenced by coastal eutrophication, as well as large parts of the open Baltic Proper and many parts of the
Bothnian coast and the West coast of Sweden. The A-value is therefore likely 7 (= more than 25% of Swedish
coastal A-areas) or 8 (= more than 50% of Swedish coastal A-areas) with Ecrit s; 2 mg 02 per litre, tile critical
limit for the survival of many important benthic animals.
• Coastal eutrophication is caused by many different types of emissions and will probably continue for several
decades. The T -value for Baltic coastal areas is difficult to estimate, but is here set to 7, which implies that the
present severe contamination is likely to last for 50-100 years.
• According to these criteria, PER for coastal eutrophication in the Baltic is very high (560).

85
1.6.4. Lake eutrophication

1.6.4.1. Effect-load-sensitivity

A derivation of the Vollenweider model is given in section 2.5. Many other load models for lake phosphorus have
been presented (Dillon and Rigler, 1974, 1975; Nichols and Dillon, 1978; Chapra and Reckhow, 1979, 1983;
Ahlgren et al., 1988). The most commonly used model todny is probably the OECD-model (r2 = 0.86), which is
graphically presented (the presuppositions and some important aspects of the model) in fig. 1.64.

Traditional mass-balance models treat lakes as reactor tanks. The basic ingredients of these classical ELS-models
nre summarized in fig. 1.65. This figure stresses the connections to catchment population and land use, the
relationships between total phosphorus (TP) and primnry production (chlorophyll), and the quantitative links to
widely used operational effect variables in water managemen~ like Secchi depth, maximum phytoplankton volume
(see fig. 1.66) and hypolimnetic oxygen demand (HOD).

It should be noted from figures 1.64 and 1.65 that the lake TP-concentration is not a biological or ecological effect
variable, it is a chemical concentration as a response of a certain lake load. In fig. 1.64 it is called a response
variable. As stressed before, the real interest is not on such chemical variables but on biological effect variables,
like chlorophyll in fig. 1.65.

Some models for lake eutrophication use the maximum phytoplankton volume (PI') as target effect variable. One
such alternative, which is based on an empirical regression between PI' and TP, is given in fig. 66. Other methods
(see Hakanson and Peters, 1995 and section 7) use more causal approaches to link TP and PI' based on the
interactions between dissolved-P, water temperature, stratification, ligh~ Secchi depth and the phytoplankton
turnover time.

Fig. 1.66 is based on data from 327 measurements from 10.0 Swedish lakes covering a broad range concerning lake
TP-concentrations (from 3 to 300Jlgll). The trophic states for the lakes included in the regression range from very
oligotrophic to hypertrophic conditions. A PI' of 5 mm3 fl is generally regarded by Swedish authorities as a
practical guideline, or a "critical" value concerning algae blooming, and 10 mm 3fl a limit for "alarm" (persson and
Olsson, 1994). Fig. 1.66 is based on TP-dnta measured from May to October. The spread around the regression line
in fig. 1.66 is, however, considerable and a TP-value of 35 llgll (typical for an eutrophic lake) can therefore (with a
95% certainty) correspond to PP-values from 0.5 (= 1O(-1.924+1.512*log(35) - 0.7» to 12 mm 311. This uncertainty
must be accounted for in practical evaluations in lake managemen~ e.g., in determining how large phosphorus
emissions a certain lake can endure.

Such a calculation, which illustrates tile practical use of tile OECD-model combined with the regression in fig.
1.66, is given in table 1.9. Note, however, that calculations like these can generally be done in a much more
relevant manner using dynamic models such as tile LEEDS-model presented in section 7.

It is evident that one can use the two operational guideline limits for the effect variable PI', tile critical PI' =5 and
tile alann PI' = 10, in tile PER-analysis. However, it is NOT evident what E-value should be given to Ecrit = PP = 5
or to Ecrit = PI' = 10. A suggestion is timt PP = 10, i.e., very high risks for frequent blooms of LOxic algae (like
"blue-greens") andior low 02-<=oncentration in the bottom water which could kill key functional groups of the
bottom fauna, may be set to E =9 (very large ecosystem effects) or to E =10 (total collapse of natural ecosystem),

86
ELS-model
Phosphorus in lakes

LAKE
TYPE
~ ~
'"< ~
Hypertrophic ~ E
~

.§. '"
u ~
Eutrophic
=80 =20
=25 ==6
Oligotrophic =2

i
EFFECT
SENSmVITY, Tw
g Ul ~ LOAD, Cin VARIABLE

Other effect variables:


Model, OECD (1982):
C = J.SS·(Cin/(l +,(rw)Y'O.82 • HOD (hypolirnnetic oxygen demand)
• Secchi depth
Legend and presuppositions: • Algal volume
River conc of tot.l-p (-TIl: Cin (in mg/m A3); range: I-ISO • FIsh community index
• Bottom fauna index
Lake water retention tjme, Tw (in years); range: O.1~lOO

Lake conc ofTI: C (in mg/mA3); range: 2.5-100 Model not applicable for:
• Monontictic lakes (model only applicable for
Degree of explanation: rA2--Q.86 (for log(C» lakes from temperate climates)
Number of lakes in the srudy: 87 • Dams/reservoirs
• Lakes with high intemalloading

Fig. 1.64. An ELS-model and ELS-diagram for phosphorus in lakes. Model presuppositions, linkages between
lake trophic indicators, lake total-P concentrations and other operationally useful measures of ecological effects
of lake eutrophication.

and PP = 5 to E = 6 (cJenr ecosystem effects) or E = 7 (substantial ecosystem effects). Such defmitions must be
given by the user of the PER-system. When this is done, the nreal (A) and temporal (T) extent of Ule given E-value
can be calculated from empirical data or from models, preferably using GIS-techniques (Geograpbical Information
System).

Models of the Vollenweider- or OECD-lype nre very useful in practical lake management, but there nre also
problems with such models. They nre, in fact, very simplistic, which implies that there may be many inherent
problems if such models ore used for predictions in individual lakes (see section 2). Most of Ulese problems nre also
emphasised in the basic publications (Vollenweider, 1968, 1976; OECD, 1982), but those caveats nre often ignored
in practical waler management (SNV, 1993a) wiUl many dire consequences, e.g., at evaluations of point source
emissions in individual lakes (see Hakanson, 1995b).

87
Population, land use
t
(WASTE SOURCE MODEL)

t
Total phosphorus loads
t
Handled by differential equations
PHOSPHORUS BUDGET MODEL

_InfI--;o>~ ~ ~>
Sediment losses and feedback

t
Total phosphorus concentration
t ~
,,--------... TOTAL PHOSPHORUS/
PHYTOPLANKTON
PHOSPHORUS/ REGRESSION
CHLOROPHYll
REGRESSION
t
t Maximum volume of
ChIrOPhYII COncentra 'on phytoplankton (PP)
/CHLQ--R-O-P-HYLL/S--EC-'-CHI....... /-CHL=:::OR;:::O;:;P;;:HYLl/P~;-;;;:RIMA~:-:-::R::Y-.....
DEPTH REGRESSION PRODUCfION
REGRESSION
t t
Secchi Depth Primary production
t
0XYGEN MODEL )

t
Hypolironetic oxygen
concemration (HOC)
Fig. 1.65. Basic elemenlS in many traditional management models for lake eutrophication (basic structure of
figure from Chapra and Reckhow, 1983).

All models of the Vollenweider-type generate uncertainties because they do not account for:

(I) Seasonal variations in phosphorus fluxes. It is very important for both TP-fluxes and phytoplanklOn productiOn
to account for seasonal variations in temperature, mixing/stratification, water clarity and load of phosphorus from
different sources. Models of the Vollenweider-type are generally derived for situations where the maximum TP-
load 1O Ille lake occurs in spring (in connection willl high tributary water discharge) and should Illerefore not be
used for, e.g., fish farm emissions which are largest during harvest in fall (Hakanson et al., 1988a).

(2) Bioavailable load of lOtal-P. The ratio between bioavailable P (= dissolved P) and total-P varies wid,in and
among lakes. This ratio is defined by Ille Kd-value, the partitioning (or distribution) coefficient. The LEEDS-model
for lake eutrophication discussed in section 7 gives a method to estimate the distribution coefficient for lake
phosphorus.

88
1 ; I
1og(PP) ~ -1.924 + 1.5t2"1o&(11')
n -= 3Z7;r 2 "" 0.76; P < 0.0001
I II /
2.0 ± 9590 con!. int '" 0.70
I..'
j'
.V
1.5 I,, I ·.···;1
.' I ../.
1.0
I
I
I
I
II.'••..•.
.....·t~/I·
...... • V' i
;0
·
I"
4It1·

...
.• . , '
..\
"
1.05
-1 OAlarm level

I..... ~
I
.. ... •• .
"-"'JJ: • '~" _s Critical level
i : .~~
' I·.··
I....
"
~~ ~, 1"'
... I 0.35
Variations depending on:
1. Temperoture
~.f • ~t·. ~•• -;. ,I
0.0 " I..• • .-;-
'-:1.' I
.1, ," .
+• • ,' •
I I, 2. Daylight
3. Water clarity
4. Biological processes
•~:k("
I
, I,
I+' .!I (pre<iation)

. I
•• • • j

.+.
..
, •• , . " -0.35 5. Empirical uncertainties

.. . . . ." I
I
,'j i (sampling, transport. storage.
-0.5 ' I
analysis. etc.)
;r:lr.~
j
' I II
~1.0
: i
• .-.s.....~
• ~ t' ,,'.
.'
II I

•• ~t "I I

-1.5
.... !
! ,, i,
t i
0.5 1.0 1.5 2.0 2.5 3.0

Oligo-
trophic
Meso-
trophic
Eu-
trophic
t Hyper-
trophic

Fig. 1.66. The relationship between total-P rIP in )lgl1; mean lake value from March to October; the figure
log(TP)

gives 10g(TP)] and maximum total volume of phytoplankton [PP in mm311; the figure gives 10g(PPlJ. Based on
unpublisbed data from E. Willen (SLU, Uppsala, Sweden). The regression line and the 95'70 confidence
intervals for the predicted y show that there exists a very strong (r2 =0.76, P < 0.0001) general relationship
between the x-variable and the y-variable for these 327 measurements from 100 Swedish lakes, but there is
also a substantial residual variation around the regression line. The 95% conf. int. for 10g(PP) of 0.7
corresponds to a PP-value of 100.7 =5.01. A lake with a summer value of TP =35 )lgl1 (typical for an
eutrophic lake) would be expected to have PP = 2.6 mm311. but the true value can thus range from 0.5 to 12.
Some of the variation around the regression line can be related to variations in lake temperature, daylight, lake
water clarity, biological predation and analytical errors. The critical PP-limit is set to 5 and the alarm limit to
10 mm 311.

(3) Internal loading ofphospborus. The models of Vollenweider-type do not account for intemalloading. This has
for many years been regarded as difficult. The LEEDS-model gives sub-models to quantify the intemalloading of
phosphorus in lakes, but many details of these processes are very complicated since they include processes like
wind/wave induced resuspension, slope processes, seiche activity, diffusion, bioturbation. etc. (Hakanson and
Jansson. 1983; Pierson and Weyhenmeyer, 1994; Weybenmeyer, 1996).

(4) Retention of phosphorus in lake water. All models of the Vollenweider-type oversimplify tile retention of
pbosphorus in lakes and they do not account for seasonal variations related to stratification and mixing (Hakanson
and Peters, 1995).

89
Table 1.9. Calculations using the OECD-model and the regression between lake TP and PP (fig. 1.66) to
illustrale how the models can be used to evaluate how high tributary load of phosphorus a given lake (data
from Lake Southern Bullaren, Sweden) can endure before reaching the critical guideline limit pp; 5 mm3 tl.

River concentration Lake concentration Max. primary production


(~ TPIl) (~TPIl) (pP in rnnn31l)

15 11.7 0.49
30 20.7 1.16
60 36.6 2.75
120 64.6 6.5 I> criticallevel=5
1.6.4.2. Area and time perspectives

About 1-5% of the total population of Swedish lakes, i.e., 800 to 4000 of 83,200 lakes, have clear changes in
ecosystem slfUcture (E; 10) due to eutrophication (SNV, 1993b). Most lakes influenced by anthropogenic
eutrophication (in contrast to naturally eutrophic lakes) are relatively large (and hence important natural
resources), shallow and located in agricultural landscapes in the southern part of the country. There are also some
lakes, like Lake Norrviken (Ahlgren, 1970), north of Stockholm, which have been heavily polluted by sewage
effluents or nutrient emissions from specific point sources.

The time perspective of lake eutrophication (the T-factor in the PER-analysis) may be illustraled with results using
the LEEDS-model (see section 7), an extensive dynamic lake ecosystem model accounting for all major processes
regulating the transpon, biouptake and effeclS of phosphorus in lakes. The example in fig. 1.67 concerns eutrophic
Lake Southern Bullaren (area 8 km 2 , mean depth 10 m). The questions in this context are: (1) How long will it take
before a new dynamic equilibrium is reached after a very heavy phosphorus load (related to an annual production of
4000 tons of rainbow trout in a fish farm in the lake; note that the pennit for fish production was for 70 tonstyr and
that the total fish farm production ,in Swedish lakes is about 3000 tonstyr in the !ale 1990s)? (2) What does the
recovery look like, a fast initial phase, or a steady, slow process? The model indicates that it would take about 5
years to reach a new dynamic equillbrium. This resullS is, of course, lake-typiCal and carmot be applied directly to
other lakes. Other lakes might react faster or slower than Lake S. Bullaren depending on differences in lake
morphometry, water turnover and waler chemistry, stratification and nutrient loading, but the recovery can oflen, in
tenns of order of magnitude, be set to about a decade.

1.6.4.3. Summary - lake eutrophication

Utilizing the criteria to rank environmental threats, as given by PER one can note that:
• Eutrophication bas altered and killed key functional organisms and changed the ecosystem slfUcture of hundreds
of lakes. This means that the E-value, as defined bere, sbould be set to 10.
• A survey presented by tlle National Swedish Environmental Protection Agency (SNV. 1993b) has shown that
800 to 4000 Swedish lakes have clear eutrophication effeclS in the fonn of alterations in key functional species or
increased amounlS of toxic algae. From tllis, the A-value may be set to 6, wbich means that ecosystems in many
regions in Sweden have been affected (Witll E; 10).

90
Fish production (tJyr)

100
500 t
.. 4000 t
• ..
500 t
.. Ot ...
A.
~

'So
I~ \ \
.:;
'"
0

~f'v...f'v...I\",1\
'.::J
«I 50
!:l
Emp.
'"tJ" J
~ .A./\
reference
0
u, ~ value ~ 34.4
'"
....
0
1 61 120.5 180 24 o
Months

B.

Alarm
level
-
Critical

A~
level
Af\ AA
AUV U !V\MM
1 61 lZO.5 180 24 o
Months

Fig. 1.67. Simulations with the LEEDS-model (see section 7) to illustrate the recovery process for (Al total-P
concentrations (fP) in lake water and (B) the target effect variable in LEEDS. maximum volume of
phytoplankton (PP) after a simulated heavy contamination. For a period 4000 tons/yr are produced in the fish
farm in Lake S. Bullaren, then production is reduced to 500 tons/yr, and fmally, the farm is closed - what do
the recovery look for TP and PP? After about 5 years, there is a new dynamic steady-state when tlle values are
lower than the critical limit. After the farm is closed down, the recovery to a new dynamic steady-state takes
also takes about 5 years .

• If the contamination is halted, most lakes recover within about 10 years. Due to political and economical
reasons, it will evidently take quite a long time to reduce anthropogenic fluxes of nutrients from land to water,
e.g .. to effectively implement agricultural measures to reduce the P-fluxes (see Svendsen and Kronvang, 1991).
The T-value in the PER-approach should preferably be related to the ecological response time if nutrient
emissions are halted, and be set according to the precautionary principle to account for inherent uncertainties. This
means that for lake eutrophication T is set to 5 (i.e., the recovery time is generally more than 10 years and less
than 20 years). This gives PER = 300.

91
1.7. Conclusions - a ranking of the threats

Table 1.10 summarises the results concerning the PER-criteria to rank different chemical threats to Swedish aquatic
ecosystems.

The most important aspect of the results in his mble is the structure to ask questions and analyse the chemical
threats. This structure is meant to be simple and useful for most types of chemical pollutnnts and for most types of
ecosystems. It is evident that very much work remains to be done in this sphere and that the results in mble 1.10
must be regarded with due reservations. The basic idea is to use the PER-criteria to minimise the element of
subjectivity and maximize the element of objectivity in these very complicated matters. The idea is to have a
general scientific framework for management to rank different threats so that time and effort can be directed to the
large problems. According to the PER-criteria, one can note that acidification is the largest chemical threat to
Swedish aquatic ecosystems, followed by coastal and lake eutrophication and mercury contamination of lakes. The
smallest problem in this survey is the radiocesium conmmination of lakes because there are no esmblished, or even
likely, effects on aquatic ecosystems from this threal The areal distribution of the cesium problem concerns not the
entire country but certain regions, and the duration of the problem could be seen in terms of a few decades rather
than centuries.

The result of the PER-ranking is:

Acidification of freshwater ecosystems (PER =700) > eutrophication of coastal ecosystems (PER =560) >
eutrophication of freshwater ecosystems (PER = 300) > mercury contamination of lakes (PER = 240) >
conmmination of organic toxins in the Baltic (PER =220) > radiocesium contamination of lakes (PER =70).

The results in mble 1.10 may he discussed from many angles and compared to other compilations of threats to aquatic
ecosystems. One such compilation is given in fig. 1.68. From the results given in this figure, one can note that lake
acidification has caused more harm for fish biology in terms of number of lakes with reduced or extinct number of fish
species than all of tile other listed threats. This agrees well with the results from the PER-analysis in mble 1.10.

Note again that the PER-approach may be used not just for Swedish aquatic ecosystems, but for most types of
ecosystems in most regions. The basic aim of the approach is to provide a general scientific structure for
"environmental diagnosis" where important elements are (1) definition of ecosystem, (2) defmition of operational
effect variables related to Ole defined chemical threat, (3) ecometric analysis, i.e., effect-load-sensitivity analysis,
(4) integration (or summation) of E or Ecrit over impact area, and (5) integration (or summation) of Ear Ecrit over
impact time.

As a comparison to Ole cases treated in olis chapter, and to demonstrate how the PER-analysis can be done also for
other siLUations, data are given in table 1.10 related to the Tesis oil spill (from Kineman et aI., 1980). 111is aCCident,
wbich is very well studied because it happened close to the marine ecology laboratory at Ask6, Sweden. On
October 26, 1977, Ole Soviet tanker Tsesis struck a rock and about 1100 tons of fuel oil were released to the Asko
coastal archipelago. The largest visible extent of the oiling area was about 34 km 2 Clear damage to zooplankton,
mussels and especially to crustaceans (but not to fish) were seen, but Ole structure of the ecosystem did not change
markedly. The PER-analysis should focus on the most sensitive key functional group, crustaceans, and for this
group the substantial damage lasted for more Olan 1 year. Note that this is just one of many accidents involving oil

92
5000

4500

4000

~
3500
Q

'".!!! 3000
'0
-
Q
.c
E 2000
2500

z"
1500

1000

500

0
c w :;;
.2 "
~ "w
C>
c
:c 3: .E 1': C>
c:
c:
.2
c:
0
c:
.!!!
~ "~ t:: "E C n; ;;;
'0
:;
.g ~
Jg
>-
>
0
0.
0
-0
0 C>
0. C
0.-
n;
g
3:
[ ~
~
0
:E
c:
Ern
"'0 ~

<l C>
..: .fi >- ".r:;
~ .2
w 0.
e c: -U "
..: 0
""
:I:
:I:
"
"0 .r:;
$
"
c:
0
c
al
c:
:;
" '5
UJ
E
0
0
0
13 8.
w
m a: 0
g"
:I:

*
a:
X
....0 E
Fig. 1.68. Different causes for the reduction andlor extinction of fISh communities in Swedish lakes. Note that
acidification in most important among these causes, but that also eutrophication has implied significant
changes in the structure of fish communities and that many other causes than chemical threats, like hydropower
constructions, ruined spawning areas and introduction of alien species, have changed the fish communities.
From EPA (1997).

spill, and that such accidents may be treated separately or taken together in the PER-analysis. This is a matter of
definition of the presuppositions. In table 1.10, the evaluation concerns just this particular accident. The PER-
number for this case-study is 56 (E =7, A =2 and T =4).
The results in table 1.10 are open to debate and improvements!

93
Table 1.10. A compilation of PER-criteria to rank the chemical threats to Swedish aquatic ecosystems
discussed in this work.

Swedish 3guatic ecos;),stems


Threat Ecosystem Erfeet vadabl' - Areal distribution Duration iii Ume PER= -PER ranking
(E) (A) (T) E*A*T priority for action
Acidification Lake 10 7 10 700 I
Mercury contamination Lake 3 8 10 240 4
Rndiocesi urn comamination Lake 2 6 6 72 6
Organic toxin contamination Marine/coastal 6 6 6 216 5
Eutrophication, phosphorus Lake 10 6 5 300 3
Eutrophication, phosphorus and nitrogen Marine/coastal 10 8 7 560 2
\0
..,.
Tsesis oil spill Marine/coastal 7 2 4 56

Effect variable: Areal distribution: Duration In time:


1=00 known or likely ecosystems effects l=no ecosystems widl E = 10 and/or E = Eerit 1= no effects
2=unlikely ecosystems effects using statistical methods 2=a few ecosystems with E = 1O!Ecrit «25 lakes or coastal areas) 2= effects (E = 10/Ecrit) for less dlan 1 month
3=likely but low ecosystems effects using statistical methods 3=several ecosystems widl E = 10/Ecrit (25-100 lakes) 3= more Ulan 1 month
4=probable ecosystems effects using statistical medlOds 4=many ecosystems with E = 1O!Ecrit (100-400) 4= more than 1 year
5=small real ecosystems effects 5=most ecosystems in a region with E = lO!Ecrit 5= more than 10 years
6;;;:clear real ccosystcms effects 6;;;:ccosystems in many regions wilh E;::; 10/Ecrit 6;::; more than 20 years
7=substantial real ecosystems effects 7=more than 25% of Swedish lakes or coastal areas with E = IO/Ecrit 7= more Umn 40 years
8=large real ecosystems effects 8=more than 50% of Swedish lakes or coastal areas widl E = IO/Ecrit 8= more than 80 years
9=very large real ecosystems effects 9=more than 75% of Swedish lakes or coastal areas widl E = 10 9= more than 160 years
10=lOtal collapse of real ecosystems IO=alllakes or coastal areas in Sweden wiUl E = lO/Ecrit 10= effects (E = 10/Ecrit) for mOre than 320 years

Note 1. For mercury contamination widl E = 3, the A- ,md T-values arc determined for the guideline limit for the target variable, Hgpi = Ecrit = 0.5 mg Hg/kg ww.
Note 2. For radiocesium with E = 2, Ule A- and T-values arc determined for the guideline limit for the target variable, cesium in fish for human consumption, Ecrit =1500 Bq/kg ww.
Note 3. For organic contamination with E = 6, A and T are determined for the "critical" limit related to EOCI-concentrations in surface sediments of 500 ~g EOCI/g org. material.
Note 4. As a comparison to the cases treated here, data are also given related to the Tesis oil spill (from Kineman et aI., 1980).

/
2. Introduction to aquatic ELS modeling
2.1. Background and aim
Aquatic ecosystems are very complex webs of physical, chemical and biological interactions (fig. 1.1).
It is generally both costly and laborious to describe their characteristics, and to predict them is even
harder. To develop scientific programs of conservation, management and remediation is an even
greater challenge. Every aquatic ecosystem is unique, and yet it is impossible to study each system in
the detail necessary for case-by-case assessment of ecological threats, and proposals for remedial
measures. In this situation, quantitative ELS models are essential for predicting, making
environmental assessments and directing intervention strategies.

2.1.1. The role of prediction


There are two important reasons to spend time and effort at quantifying important environmental
features and processes. Formulating quantitative descriptions aims at representing aquatic ecosystems
the way they appear, and as closely to the "truth" as possible. Quantitative prediction is necessary for
testing assumed scientific relationships; whether a change in one ecosystem variable may affect
another variable. Thus, predictions obviously have great importance for environmental management.
Environmental managers, policymakers, and the general public often have a considerable interest in
knowing what the probable environmental effect will be of a certain policy or measure, such as
liming or building sewage treatment plants. The ability to predict important goal variables, such as pH,
the Secchi depth or the algal bloom intensity (as measured by Chl) in aquatic systems is thus very
useful in practice as a communication tool between natural scientists and the rest of the world.
Furthermore, high predictive power is an important scientific goal in itself. The certainty with which
we can predict changes in water quality from changes in external factors, such as nutrient loadings, is a
direct, quantitative indicator of how well we understand scientific relationships (Peters, 1991).
This certainty is commonly referred to as the predictive power, or the forecasting power.

An alternative approach to predictive models is the use of conceptual models, which is still attractive
among some environmental analysts. A conceptual model can be a scheme over important features in
an ecosystem, such as fish, nutrient concentrations, etc., with arrows indicating the causal paths in the
ecosystem; i. e., to indicate how one ecosystem feature is influenced by others. However, a conceptual
model does not convey any quantitative information about what happens to the ecosystem if external
factors (climate, pollutants, large-scale fishing, etc.) change compared to present conditions. Thus, the
complete structure of a conceptual model is often difficult to test against empirical data, because there
is no direct connection between a conceptual model and empirical indicators. Instead, its
credibility can to some extent be established indirectly (indirectly tested) by empirically testing
measurable theories derived from the model. However, within the field of lake restoration, the
success record of conceptual models has been relatively meager (Peters, 1991). Conversely,
quantitative, predictive methods have thus far been instrumental in massive and successful
programs of eutrophication and acidification control in many lakes around the world. Thus, ELS-
analysis of aquatic systems relies on predictive modeling because of its greater scientific potential.

2.1.2. An unambiguous definition of scientific method


If one agrees that science has a unique and important role in society, and that it is not a complete waste
of resources to spend large sums of governmental and private money on scientific research, then this
insight should lead one to conclude that the unique status of science requires that it be defined in an

95
unambiguous manner. There are many variants available in the philosophy of science which in one
way or another defines science as "something that scientists do" (Peters, 1991). According to such
definitions, we would be able to consider astrology, alchemy and various holy scripts as science, since
many scientists have been involved in those fields during history. However, if we admit that these
fields have generated very limited scientific success and predictive power over the years compared to
the primary parts of, e. g., physics, medicine and ecology, then we apparently need a different
definition of science.

To this date, Karl Popper's (1902-1994) widely used demarcation criteria is the only available method
which unambiguously separates alchemy and religion from the gravitation theories in physics.
These criteria state that (1) the hypothesis or theory has to be supported by some kind of observation
and (2) that the hypothesis or theory is refutable (or testable); i. e., that it can be falsified with
evidence of the opposite. Using these criteria, hypotheses and theories are repeatedly tested,
completely or partially refuted, and improved, thus developing our knowledge in various scientific
fields and bringing it closer and closer to the unattainable goal; the truth. Constructs which do not
meet Popper's requirements are referred to as metaphysical, or pseudo-scientific. Research fields
where irrefutable constructs play a central role may produce excellent descriptions of important and
relevant features in their area, although the predictive power regarding important goal variables (and
thus the quantitative understanding of scientific relationships) may nevertheless be very poor (Peters,
1991). Yet, metaphysical constructs may very well inspire and promote the development of testable
hypotheses, theories and models with high predictive power - although it is important to bear in mind
that metaphysical constructs do not have a scientific value of their own (Peters, 1991).

2.1.3. Testable predictive models


Predictive power and refutability are crucial components of practically useful ELS models. Such
models must satisfy some categorical features that make them simple and reliable tools for
environmental management:

- they must be characterized by a relevant and simple structure, i. e., involve the smallest possible
number of driving (input) variables;
- the values of the necessary driving variables should be easy to access and/or to measure;
- the models must be validated for many different aquatic ecosystems across a wide range of
environmental characteristics (regarding ecosystem size, latitude, climate, maximum water depth,
etc.)

In broad terms, the parameters used in environmental models may be divided in two categories:

1) variables for which site-specific data are easily available, such as lake volume, mean depth, water
discharge, amount of suspended particulate matter in water, etc.;

2) model constants for which generic (= general) values are used due to the lack of easily measurable
site-specific data, e. g., the sedimentation rate of particles and/or rates for internal loading of matter
from the sediments.

The variables belonging to the first category are often called "site-specific variables", or
"environmental variables" or "lake-specific variables". They can generally be measured relatively

96
easily and their experimental uncertainty should not significantly affect the overall uncertainty of the
model predictions of the target variable(s).

The second category, the "model constants", is sometimes referred to as "calibration constants". They
are often difficult to empirically access for each specific system, such as the transfer rates from the
sediment to the water, the deposition velocity of X from water to sediments, the migration rate from
catchment to lake, etc. The model constants may contribute significantly to the model uncertainty, so
their values must be established from extensive, critical tests against data from surveys of many
different aquatic systems.

It may be mentioned that it is rather common among some environmental modelers to use models
whose model constants are not constant but calibrated or tuned differently for different sites. Such
practice is, however, very risky because it can support untestable (irrefutable) model structures. A
poor model can be calibrated to give good results at one site, and then re-calibrated to give equally
good results at another site - but such results are prone to being examples of "the right answers for the
wrong reason" (Peters, 1991). Hence, predictions from site-specifically tuned models should be
regarded as unreliable and ill-suited for ELS-analysis.

Critically validated widely applicable models with high predictive power are difficult to develop.
Many generally valid model constants in such models are (see Monte et al., 1997) defined from
"collective parameters" (see fig. 2.1.). In many circumstances, the values of such important driving
collective parameters integrate many compensatory effects (see fig. 2.1) of the different phenomena
occurring in a complex ecosystem where "everything depends on everything else". Examples of such
collective parameters in the freshwater environment are the "migration velocities" of the metal or
radionuclide from water to sediments, the "effective removal" rates and the "soil permeability
coefficient" of a radionuclide from the catchment to a water body (Monte, 1995).

Models based on such "collective parameters" show a unique and important feature: their predictions
have a relatively low uncertainty despite the large range of the environmental characteristics and the
lack of site-specific values of the model variables. The main lesson is that in predictive modeling, it is
seldom necessary, or wise, to account for "everything". The difficult task is to omit processes which
may add more uncertainty than predictive power to the given target variable.

The traditional modeling philosophy is, indeed, based on what has been called a "bottom-up" or an
"assembly line" or a "pyramidal" structuring of the set of the occurring processes (see fig. 2.2). It is
assumed that some fundamental processes, belonging at the top vertex of the logical pyramid, may be
modeled in terms of logical-mathematical primary principles from which all other natural processes
may be derived.

One can also call this structure "Euclidean-like" since the first model of this kind was developed by
the ancient Greek mathematician Euclides. Euclidean geometry is indeed the mathematical model of
the physical space. This classical modeling approach is based on the principle that the knowledge of
nature may be derived from the principles of such primary models.

97
Fig. 2.1. Illustration of the concept "collective variable". The figure shows an integration process where a target
substance and/or animal and/or effect variable (y) has a certain pathway (from 0 to 15 according to the given
isolines) in the given ecosystem where the x-variables can influence y. The frequency distribution illustrates the
variability for x and the characteristics value (=median). "Minus-effects" will be balanced by "plus-effects" and
the characteristic x value may be used to best describe and predict how x influences y.

Environmental models based on collective parameters are structured differently, more similar to a
web than to a pyramid. Each process is indeed related to a variety of other phenomena and there is
no reason to use few of them as fundamental starting points for understanding and predicting all the
others. If it is possible to find processes that may be modeled by means of mathematical formulae
based on collective parameters, this approach may be tested - the approach which yields the highest
predictive power should be the preferred one. A variety of past experiences (see Peters, 1991;
Håkanson and Peters, 1995) demonstrate that complex models based on general principles are often
more uncertain, and yield less predictive success than simpler models based on collective parameters.

In a strict sense, there is no such thing as a general (= generic) ecosystem model, which works equally
well for all ecosystems (of a give type) because all models need to be tested against reliable,
independent empirical data and the data used in such validations must of necessity belong to a given

98
restricted domain. If this domain is equal to the entire population of ecosystems of the given type, then
and only then, is the model generic in the strict sense. The complexities of natural ecosystem always
exceed the complexity and size of any model. Simplifications are always needed, and this entails
problems. There are dynamic mass-balance models available which have been tested over such wide
ranges (for, e. g., radiocesium) that it is tempting to label them generic, but there will always be an
ecosystem with properties outside the given domain for which the model would yield poor predictions.
This is why modeling can be pictured as a two-sided coin: One needs the equations as well as the
range where the equations apply.

Fig. 2.2. Schematical illustration of hierarchical modes of thinking in predictive modeling. The target y-variables
in this example are radiocesium in fish eaten by man and Cs-concentration in lake water (from Håkanson ,
1997).

The aim of this section is to present some fundamental structural components for some of the ELS
models discussed in the first part of this book. The next section will give different types of more
comprehensive models based on the modeling approach using collective parameters. There are three
specific goals in this section:

1. To present the basic components of a dynamic mass-balance model for lakes using differential
equations. This is the backbone of the famous Vollenweider (1968) model and many following
models. If these basic elements are properly understood for one element for one type of ecosystem
(like phosphorus for lakes), then the same approach can be used in all analogous contexts. Since basic
mass-balance models of the Vollenweider-type are very simple and do not account for many important
processes, there are also strong reasons to discuss models which account for (and predict), at least,
some additional processes to inflow, outflow and net sedimentation, like internal loading (substance
fluxes from the bottom sediments) and seasonal variations. The aim is to give a technical account of

99
methods to build models of the ELS-type, which are used in practical contexts, like in the PER-
analysis, but the aim is NOT to give a thorough compilation of methods to test models. Such reviews
of testing methods can be found in Håkanson and Peters (1995).

2. To present the basic elements of empirical/statistical ELS modeling. The example given here
uses nutrients in coastal areas, and the aim is to go through the steps behind the ELS model presented
in fig. 1.60 (using the oxygen saturation in the deep water as the operational effect variable for
coastal eutrophication). If the steps in this derivation are understood, they can be applied in many
similar contexts.

3. To discuss some important concepts of ELS models, namely time and area compatibility of data,
the ecometric matrix and some statistical aspects related to regressions, and a section on sensitivity
and uncertainty analysis, two fundamental principles of model testing (see Hinton. 1993; Hamby,
1995; IAEA, 1998).

The intention here is NOT to write a literature review ("who did what") but to cover fundamental
structural components of ELS models ("how it works", "how to structure aquatic ecosystems for
predictive models", etc.). Only a few key literature references are given in the text.

Many factors may influence how an effect variable varies among aquatic ecosystems. The ELS
analysis aims at identifying the most important factors in this respect. Frequently, there are no causal
explanations of phenomena that can be established statistically beyond dispute. One of the
advantages of the empirical/statistical approach is that it provides a possibility to rank factors
exerting influence on an effect variable so that future research can be concentrated on these factors,
Naturally, when using models for the ecosystem level (for entire lakes, coastal areas, etc.), it is not
possible to describe phenomena at the individual, organ or cell levels. All methods have their
limitations.

The following sections will present different types of ELS models. All these models, however, have
three common features: They all aim to predict a defined target effect variable; they are all based on
operationally defined load and effect variables; and they are all meant to be practically useful in lake
management e. g., to simulate effects from remedial strategies. On the other hand, they are all different
in the following principal ways:

1. There is a classical modeling approach (see fig. 1.66) where a simple dynamic mass-balance
approach is used to calculate a concentration for a chemical, which is related to a target effect variable
by means of a regression. This is the approach presented in section 2.7, the Vollenweider- and OECD
models for lake TP concentration (see fig. 1.65) and the regression between TP concentration in lake
water and maximum volume of phytoplankton (PP: see fig. 1.67).

2. There is a more comprehensive general dynamic model structure available for some substances (e.
g., radiocesium and phosphorus). In addition, this structure may in the future be applied to more
substances, such as dioxins, PCB or nitrogen. This model structure includes many of the components
illustrated in fig. 2.3.

100
Fig. 2.3. The LEEDS model. A general, dynamic model for phosphorus in lakes, including a phosphorus budget
for Lake S. Bullaren (with a fish farm producing 500 tons/year). Arrows indicate phosphorus fluxes while boxes
indicate phosphorus masses. From Håkanson (1999).

3. There is an empirical (=statistical) regression model structure. The derivation of this model uses
specific steps to structure the empirical data before the statistical analysis (which is stepwise multiple
regression analysis in this example). One example is the model for the oxygen saturation in deep water
of coastal areas (O2Sat), which is presented in section 2.7.

101
There are also various mixes of these three structures available, such as a dynamic model structure
based on an empirical model, and a dynamic model structure transformed from an empirical
model. These are not covered here but are exemplified in Håkanson (1999).

Brief summary
- ELS models should be general and testable.
- High predictive power is instrumental to successful modeling.
- Predictive models may be very useful tools for deciding how to restore damaged aquatic ecosystems.
- Important steps in the development of dynamic models are calibration, validation, uncertainty
analysis and sensitivity analysis.

2.2. Ecosystem sensitivity

The difference between aquatic ecosystems in sensitivity to pollutants and the importance of
sensitivity was repeatedly stressed in the first chapter of this book. This section will go more into
detail as to what determines the sensitivity of lakes and coastal areas. We will start with coastal areas
since there is a special and outstandingly influential factor that affects the sensitivity of coastal
ecosystems, namely conditions in the outside sea. This factor is of course not relevant at all for lakes,
while many other sensitivity factors are quantified in a corresponding way for both lakes and
coastal areas.

2.2.1. Basic hydrodynamic principles and processes for coastal areas


A coastal area may de defined and characterized in many ways, for example, according to territorial
boundaries, pollution status, vertical temperature-based or salinity-based stratification
(thermoclines/haloclines), etc. One fundamental and very broad way of characterizing the entire
system is according to physical geographical zonation into:

1. The drainage area; also called the catchment area or, in American literature, the watershed. The rain
falling on this area will, in due course, find its way to the open water areas. The drainage area of the
Baltic Sea covers 1,700,000 km2, which is more than 4 times larger than the entire water area (415,266
km2). Due to the climatological and geographical differences between the catchment areas of the
different rivers, the water transport (and the chemical characteristics of the water) is very different in
different rivers. There are significant seasonal variations in the river discharge. The maximum runoff
generally occurs in the spring during the thawing period.

2. The coastal zone; the zone inside the outer islands of the archipelago and/or inside barrier islands.
This is the zone in focus in this section, and for the following ELS models. The retention time of the
water and the characteristics of the different types of pollutants may vary significantly between coastal
areas. The coastal zone is of special importance for recreation, fishing, water planning and shipping
and a zone where different conflicts and demands meet. The natural processes (water transport, flux of
material and energy and bioproduction) in this zone are of utmost importance for the entire sea. It
often referred to as the "pantry and a nursery" for fish, shellfish and other marine organisms.

3. The transition zone; the zone between the coastal zone and the deep water areas. This is by
definition the zone down to depths where episodes of resuspension of fine material occur in
connection with storm events and/or current activities (at about 50 m water depth in the Baltic; see fig.

102
1.49). The conditions in terms of water dynamics, distribution of pollutants (like nutrients, metals and
chlorinated organics), suspended and dissolved materials in this zone are of great importance for the
ecological status of the entire system. This zone dominates geographically the open water areas
outside the coastal zone in the Baltic.

4. The deep-water zone; by definition the areas beneath the wave base. In these areas, there is a
continuous deposition of fine materials. It is the "end station" for many types of pollutants and these
are the areas where conditions with low oxygen concentrations are most likely to occur.

Many factors influence the water exchange in coastal areas (fig. 2.4). Emissions of nutrients or toxins
from point sources cannot be calculated into concentrations without knowledge of the water retention
time. If concentrations cannot be predicted, it is also practically impossible to predict the related
ecological effects. Thus, it is important to introduce some basic concepts concerning the turnover of
water in coastal areas.

Fig. 2.4. Schematic illustration of key processes regulating water exchange in coastal areas (from Håkanson et
al., 1986).

The water exchange varies in time and space in any given coastal area. It can be driven by many
processes, which also vary in time and space. The importance of the various processes will vary with
the topographical characteristics of the coast, which do not vary in time, but vary widely between
different coasts. The water exchange sets the framework for the entire biotic life; the prerequisites
for life are quite different in coastal waters where the characteristic retention time varies from hours to
weeks. The water retention is also a direct determinant of how sensitive a certain coastal area is to
local pollutant emissions compared to the influence from the open sea waters outside.

Factors influencing the water exchange are:

103
The freshwater discharge (Q; given in volume per time unit) is the amount of water entering the coast
from tributaries and groundwater per time unit. In small bays with large tributaries (estuaries), the
Q-factor may be the most important factor for the water retention time.

Tides. When the tidal variation is larger than about 40 cm, it is a key factor for the surface water
retention time. The tidal range is only about 3 cm in the southern Baltic Sea.

Water level fluctuations always cause a flux of water. These variations may be measured with simple
gauges. They vary with the season of the year and are important for the water retention time of
shallow coastal areas. Thus, the mean depth is a useful coastal parameter.

Layer boundary fluctuations. Fluctuations in the thermocline (temperature boundary) and the halocline
(salinity boundary) may be very important for surface and deep water retention times, especially along
deep and open coastal waters.

Local winds may create water exchange in all coastal areas, especially in comparatively small and
shallow coastal sections.

Thermal effects. Heating and cooling, e. g., during warm summer days and nights, may give rise to
water level fluctuations which may increase the water exchange. This is especially true in shallow
coastal waters since water level variations are particularly linked to temperature alterations in such
areas.

Coastal currents (see fig. 1.48) are large, often geographically concentrated, shore-parallel movements
in the sea close to the coast. They may have an impact on the water retention time, especially in coasts
with a large topographical openness.

The theoretical water retention time (Tw) for a coastal area is the time it would take to fill a coast of
volume (V) if the water input from rivers is given by Q and the net water input from the sea is R., so
that Tw = V/(Q + R). This relationship does not account for the fact that the real water exchange
normally varies temporally, areally and vertically.

There are several methods of determining or estimating the water exchange (see Håkanson et al.,
1984):

- The freshwater input to a bay may be used as a "tracer", and the salinity or the conductivity of the
water may be used to calculate the water exchange by means of mass-balance equations.

- Instead of using the freshwater as a tracer, one may also use real tracers, like dye tracers (e. g.,
rhodamine, a red dye). The dye tracer method requires quite a lot of special equipment and trained
staff.

- The direction and velocity of water currents may be measured quite simply with inexpensive so-
called current meters, which automatically measure the mean direction of the flow and water velocity
for the period of registration. If several current meters are placed in a given section area, the water
exchange can be determined for the coastal area

104
- The water level may be measured by different types of gauges (permanent - moveable, continuously
recording - manually handled). The water exchange can subsequently be determined from gauge data
(differences in water level with time) if the area and volume of the coast are known.

One fundamental abiotic factor that, together with the morphometry (i. e., the size and shape), sets the
framework for the marine organisms is the salinity (fig. 1.1). The salinity in the open water areas
outside the Swedish coastal zone varies from about 2-4 ‰ (=psu) in the Bothnian Bay, via 4-6 ‰ in
the Bothnian Sea and 6-8 ‰ in the Baltic Proper, to values in the range of 20-30 ‰ in the Kattegat and
Skagerrak. The deep water at any site is always more saline than the surface water. Between 50 and 70
m there is a fairly rapid increase in salinity. This steep inclination in salinity is called the halocline.
Beneath it one finds significantly saltier and denser water masses.

During summer, there is often a zone with a steep gradient in temperature in and outside coastal
waters. This gradient is called the thermocline and the water above the thermocline is often referred to
as the surface water and the water beneath the thermocline as the deep water or the bottom water.
This means that in late summer, one finds warmer, less saline water on top of colder water with
approximately the same salinity. During winter, the temperature increases steadily from about 2°C at
the surface to about 4°C near the bottom.

These two boundary layers, the thermocline and the halocline, are acting as interfaces for the transport
of water and pollutants carried by the water. In the coastal zone, the thermocline is generally at a water
depth of about 10 m (Persson et al., 1994; see fig. 2.5 for illustration). The theoretical deep water
retention time is generally longer than that of the surface water, and the deep water is often
exchanged episodically. The mixing between surface water and deep water is generally very limited
during stratified conditions, but extensive during homothermal conditions (see Persson and Håkanson,
1996).

Fig. 2.5. Illustration of thermal


stratification of a coastal area
(from Carlsson et al., 1998).

2.2.2. Fundamental sedimentological principles and processes for coastal areas


As a background for the following modeling sections, we will also give a brief discussion on some
fundamental principles and processes for coastal sedimentology.

The sediments reflect what is happening in the water mass and on the bottom - they may be
regarded as a tape recorder of the historical development and are often called "the geological
archive". The sediments also affect the conditions in the water via, e. g., resuspension and diffusion
processes and by the fact that the animals living in the sediments play a fundamental role in the
ecosystem. By extracting sediment cores and conducting a number of analyses, information is
obtained on changes that have taken place in the ecosystem (see Jonsson, 1992).

105
The grain size and/or the composition of the material are often used as criteria to distinguish different
sediment types. Alternatively, one can differentiate between different sediment types by means of
functional criteria (like erosion, transportation and accumulation) of coarse sediments (friction
material) or fine sediments (cohesive material).

Thus, regarding bottom dynamic conditions (erosion, transportation and accumulation), we will use
the following definitions (from Håkanson, 1977):

Areas of erosion (E) prevail where there is no apparent deposition of fine materials but rather a
removal of such materials, e. g., in land uplift areas (see fig. 2.6) or on steep slopes (E-areas are
generally hard and consist of sand, gravel, consolidated clays and/or rocks).

Fig. 2.6. Present-day land uplift


in the Baltic Sea region. Values
in mm/yr. From Voipio (1981).

Areas of transportation (T) prevail where fine materials are deposited periodically (areas of mixed
sediments). This bottom type dominates the open parts of the Baltic, where wind/wave action regulates
the prevailing bottom dynamic conditions (see fig. 2.7). It is sometimes difficult in practice to separate
areas of erosion from areas of transportation.

106
Areas of accumulation (A) prevail where the fine materials are deposited continuously (soft
bottom areas). These are the areas (the "end stations") where high concentrations of pollutants may
appear.

Fig. 2.7. The ETA-diagam giving


the relationship between the
effective fetch (the free water
surface over which winds
influence waves), the water depth
and the potential bottom dynamic
conditions. DE/T is the water depth
separating E- and T-areas. DE/T
can be predicted from the given
equation.

Geochemically, fine sediments behave differently as compared to coarse materials. From the basic
Stokes' equation for settling particles (see fig. 2.46 later), as well as for convenience, the limit between
coarse and fine materials can be set at a particle size of medium silt (0.06 mm).

The generally hard or sandy sediments within the areas of erosion and transport often have a low
water content, low organic content, and low concentrations of nutrients and pollutants. The
conditions within the T-areas are, for natural reasons, variable, especially for the most mobile
substances, like phosphorus, manganese and iron, which may react rapidly to alterations in the
chemical "climate" (given by the redox potential) of the sediments. Fine materials may be deposited
for long periods during stagnant weather conditions. In connection with a storm or a mass movement
on a slope, this material may be resuspended and transported up and away, generally in the direction
towards the A-areas in the deeper parts, where continuous deposition occurs.

It should be stressed that fine materials are rarely deposited as a result of simple vertical settling in
natural aquatic environments. They move to a much greater extent sideways - the horizontal velocity
component is at least 10 times larger, sometimes up to 10,000 times larger, than the vertical
component for fine materials or flocs which settle according to Stokes's law (see Bloesch and Burns,
1980; Bloesch and Uehlinger, 1986).

Thousand-year-old sediments influence the Baltic ecosystem today. When the old bottom rises after
having been depressed by the glacial ice (fig. 2.6), they will eventually reach the wave base, which is
the water depth above which the waves can resuspend the sediments. The wave base depends on the
effective fetch, and the duration and velocity of the wind; during storms, it may reach water depths of
about 50 meters in the Baltic Proper (outside the coastal zone). So, as a result of land elevation, the
old sediments deposited hundreds and thousands of years ago will be resuspended. In this way,
the carbon, nitrogen and phosphorus contained by these sediments, as well as metals and mineral

107
particles, will again enter the ecosystem of the Baltic, perhaps thousands of years after they were
originally deposited onto the bottom in a considerably calmer environment Sediment surveys have
demonstrated that as much as 80 % of the material sedimenting onto the deep bottoms may be old
eroded material (see Jonsson, 1992). When nutrient fluxes from land uplift to the Baltic Sea waters
have been accounted for, the nutrient mass-balance there has been substantially revised - land uplift
apparently influence nutrient concentrations much more than nutrient fluxes from land.

2.2.3. A coastal sensitivity index (SI)


The sensitivity of coastal areas can be expressed in quantitative terms, using the Sensitivity Index (SI)
for coastal areas, which was developed by Håkanson and Bryhn (2008). This index is valid for coastal
areas with a marginal influence from tides, which, for instance, is the case along the coast of the
Baltic Sea.

The exposure (Ex, or the topographical openness) was defined in fig. 1.6 as:

Ex = 100 · At / Ab (2.1)

where At is the section area (in m2, see fig. 2.8 for definition) and Ab = the enclosed coastal surface
area (m2).

Fig. 2.8. Illustration and definition of


the section area and the topographical
openness or exposure, which are used
to define a coastal area.

The definition process includes drawing the borderlines so that Ex attains a minimum value (see fig.
1.6). The Ex value of the defined coastal areas is of fundamental importance to the sensitivity with
respect to pollutants, and Ex is therefore included in the sensitivity index.

Furthermore, the relative prevalence of ET bottom areas determines the upward transport of pollutants
from sediments and can be derived from another morphometric parameter, the dynamic ratio (DR).
DR is, in turn, calculated from the surface area and the mean depth:

DR = √Ab/Dm (2.2)

The relationship between DR and the relative prevalence of ET areas is illustrated in fig. 2.9. At a
threshold value of DR=0.25, ET areas are particularly scarce. At DR<0.25, the sediment transport on
ET areas is dominated by slope processes, which means that particles due to gravity are falling from

108
steep ET bottoms to adjacent waters. At DR>0.25, ET areas are instead to a much greater extent
influenced by winds and waves, and particle are thereby resuspended from sediments by these forces.

Fig. 2.9. The relation between the dynamic ratio (DR) and the proportion of bottom areas dominated by erosion
and transport processes (ET). The threshold value of 0.25 is used in the following sensitivity index to separate
deep areas from shallow coastal areas. From Håkanson and Bryhn (2008).

Thus, very enclosed areas, areas with very steep shores and very shallow coastal areas should be most
sensitive to nutrient loading. This is expressed in a simple manner in by eq. 2.3, which defines the
sensitivity index (SI, dimensionless) from Ex and DR.

If DR ≥ 0.25, then SI = √((DR/0.25)/Ex) and


If DR < 0.25, then SI = √((0.25/DR)/Ex) (2.3)

DR generally varies between 0.06 and 6 and Ex varies between 0.002 and 1. By accounting for the
definition of SI (eq. 2.3), SI will generally vary between 0 (not sensitive) and 100 (extremely sensitive
areas). The class limits of SI are given in table 2.1.

Table 2.1. Sensitivity of coastal areas based on the sensitivity index (SI)
Sensitivity class SI
Not sensitive 0-1
Moderately sensitive 1-5
Sensitive 5-10
Very sensitive 10-50
Extremely sensitive >50

In a dataset of 478 Baltic Sea coastal areas (see Håkanson and Bryhn, 2008) regarding surface area,
section area and mean depth, the sensitivity index varied between 0.69 (not sensitive) to 116
(extremely sensitive); the frequency distribution was positively skewed and the mean value was higher
(5.7; sensitive) than the median value (3.8; moderately sensitive). In this dataset, there were 2 (0.4%)
"extremely sensitive" coastal areas, 50 (10.5%) "very sensitive" coastal areas, 121 (25.3%) "sensitive"

109
coastal areas, 301 (63.0%) "moderately sensitive" coastal areas and 4 (0.8%) "not sensitive" coastal
areas according to the sensitivity classes in table 2.1. The class limits and the categories in this table
could and should, of course, be discussed.

However, the definition of the sensitivity index is well motivated by empirical data and process-based
dynamic modeling. Since the exposure and the dynamic ratio are easy to define and understand, SI is
also easy to apply in practice in coastal management. Using geographical information systems (GIS)
based on digitized bathymetric data, the SI may rather easily be calculated for any given coastal area.

2.2.4. Lake sensitivity


Lake surveys also display a wide range with respect to sensitivity. The exposure (Ex) is irrelevant for
lakes since there is by definition no substantial water inflow from the sea to lakes. However, the water
retention time, as well as the dynamic ratio and other parameters that regulate sediment processes, all
determine the sensitivity of both coastal and lake ecosystem responses to anthropogenic activities. At
present, there is no sensitivity index available for lakes, but several examples of varying extents of
sensitivity have already been discussed in this book. Some of them are:

- Bedrock and soil type determine the sensitivity to lake acidification.


- pH, trophic state and morphometry influence the sensitivity to mercury pollution.
- Conductivity and water residence time affect the sensitivity to radiocesium pollution.

Furthermore, the water residence time influences the sensitivity to eutrophication (section 1.6). More
sensitivity factors regarding lake eutrophication will be elaborated in further detail in coming sections.

Brief summary:
- Water inflow and outflow are very important for the sensitivity to pollutants.
- Another influential factor is the relative distribution of bottom sediment types.
- The sensitivity of coastal waters can be quantified and ranked using the Sensitivity Index (SI).
- Conditions in the outside sea are often decisive for the water quality in coastal areas.

2.3. Time and area compatibility of data

Time and area resolution are fundamental concepts in the ELS-analysis. As an example, the time
resolution of the effect variable Hgpi will first be examined (fig. 2.10). Hgpi is the Hg-content in 1-kg
pike caught during the spawning period, generally in March/April. The value depends on how the pike
has lived and what it has eaten during a fairly long period before being caught The Hg-concentration
in pike is an integrated value for the entire environment of the pike and its prey (see fig. 2.1). It
may be said that the Hg-content in pike has a certain halflife.

If the Hg-contamination stops, it would take about 3 years before the Hg-content in 1-kg pike
decreases to half its original level (Lindqvist et al., 1991). Note that there are different fish each year
in the category defined as 1-kg pike. Thus, the Hg-content in pike depends on the Hg-load supplied to
the entire environment and not only to the clump of reeds where the pike was caught, as well as to
the biological, chemical and physical conditions in this environment over a long period since these
conditions influence the distribution of the Hg-load on different carrier particles and the bioavailability
of the Hg-load. The pH of the water is important for the binding of mercury to different types of

110
carrier particles and for how Hg is distributed among different Hg-forms, such as Hg0, Hg+ and
methyl-Hg (see section 1.4.1.).

Fig. 2.10. Illustration


of time compatible
data using the Hg-
content in 1-kg pike as
an example. From
Håkanson (1990).

It is not, however, the pH of the water when the fish is caught that is of interest but the pH of the
water during a long previous period. One must, thus, look for a mean value or a corresponding
value for the pH-level which applies for a period before the fish is caught and which is comparable
with the time taken to accumulate a given Hg-content in the fish. Since lake pH varies much within
lakes at a given time and with season of the year, a mean, median or characteristic long-time lake pH
value may also be seen as a "collective" variable. The pike and its prey will move around in the lake
and this will cause an integrating effect so that the pH at a certain site at a certain time will have a
relatively little predictive power for Hgpi. It is the integrated effect that regulates how the
characteristic lake pH influences the actual Hg-concentration in fish.

In the same way, one must know which area resolution applies to the effect variable in question. The
pike is a stationary predator, but the fish eaten by the pike and the food of such fish come from a
much wider area than the area around the home territory of the pike (fig. 2.11). The biological
contact area for one pike is frequently the entire lake, or at least an area of several tens of km2. The
load and sensitivity variables which can explain why a certain effect variable attains a certain value
must then emanate from this entire area.

All effect, load and sensitivity variables must therefore be area and time compatible in the same
way as Hgpi and pH. It is important in these ELS-contexts to understand that one must often use
statistical methods which average the spatial and temporal variability (see fig. 2.1). Each effect, load
or sensitivity variable has a certain distribution within the ecosystem during the defined period of time

111
and in this connection one must start from representative mean or median values from such
statistical frequency distributions. A very important requirement is to demonstrate that representative
mean (or median) values are available. There is nothing remarkable in this but it requires the
preparation of a sampling strategy on the basis of these conditions.

Fig. 2.11. Illustration of area


compatible data using the Hg-
content in 1-kg pike as an
example. From Håkanson
(1990).

Consider again the Hg-example, when one catches the pike at a certain time of the year (generally in
connection with the spring spawning) in order to catch enough fish and to create standards for time
variations. One uses fish which are as close to 1 kg as possible from as many places as possible within
the lake. Then one determines an area-typical value for this operational effect variable by means of
regression between the Hg-concentration in fish and the fish weight and, finally, standardizes the Hg-
concentration to the weight of 1 kg (see fig. 2.12).

Fig. 2.12. Regression between


empirical data on fish weight
(g ww) and Hg-concentration
in fish muscle (mg Hg/kg ww)
for 18 pikes caught in April
and May 1987 in lake 2201
(Selasjön, Sweden). The r2
value is 0.68, which is a rather
typical value for the
relationship between Hgpi and
fish weight The requested
Hgpi value for 1-kg pike is
1.94 in this case.

112
The certainty of the regression, or the statistical correlation between the variables in fig. 2.12, is
given by the r2 value, which can be between 0 and 1 and describes the spread of data pairs (the
circles in this figure) in relation to the regression line. An r2 value of 1.00 means that all data pairs
are located on the regression line, while r2=0.00 means that data pairs are scattered like a bee-swarm.
In fig. 2.12, the r2 value is 0.68, which means that the circles are fairly, but not very, close to the
regression line. The r2 value will be discussed in more detail in coming sections.

Mercury concentrations in water are generally low and expensive to determine. In order to establish
the Hg-load one could, in the ideal situation, place a number of sediment traps at different sites in the
lake for several years and analyze the material collected in the traps for Hg (fig. 2.13). This may be
measured as total-Hg or different fractions of Hg, such as methyl-Hg (Lindqvist et al., 1991).
Naturally, very little of the mercury found in the sediment traps will enter the fish. The mean Hg-
content from the sediment traps will provide an integrated and indirect measure of the Hg-load to
the lake for the registration period.

Fig. 2.13. Schematic illustration of how time and area compatible data for effect, load and sensitivity variables
may be obtained from field investigations in natural lakes using sediment traps. From Håkanson (1990).

Most water chemical and sediment variables vary with sampling location and time within a lake. One
can ask: Which variant of such a variable should be chosen in a model? Fig. 2.14 presents the basic
issue: How to establish the most time-compatible variants of lake variables. There are four x-
variables in fig. 2.14 and one y-variable. All these variables appear with different seasonal patterns.
There are small-scale (daily-weekly) and large-scale (monthly-seasonal) variations. Assume that y
depends on these x-variables, in the same way that Hgpi depends on Hg-load and pH or Secchi depth
depends on color and total-P (see Håkanson and Peters, 1995). Which variants of the variables
would be most compatible with, say, monthly mean Secchi depth? From fig. 2.14, one can note that
the highest r2 value exists between y and x3 (r2 = 0.99), while the lowest value is found between y and
xl (r2 = 0.83).

We have already stressed the concept of time compatibility as very important in ecosystem studies
and ELS modeling. Some substances are conservative in the sense that they rarely take part in lake
processes (like chloride), but most substances are reactive in the sense that they may be altered in the
system by physical, chemical and biological processes. The reaction time may differ very much
between substances and different forms of the same substance. Because of such complicated
interdependencies, it can be very difficult to make predictions at the ecosystem level.

113
Fig. 2.14. Principal illustration of time compatibility. From Håkanson and Peters (1995).

The problem is illustrated with real data in fig. 2.15. Lake total-P affects the production of plankton
and algae, and hence also water clarity and Secchi depth (see table 1.6). If one uses the long-term
mean Secchi depth (Sec36, the mean value from 36 months, in m) as a y-variable, one can ask which
variant of total-P should be used to produce the predictive model with the highest r2 value. Fig. 2.15
tries to answer that question. There are significant differences among the different variants - from r2 =
0.02 to 0.47! The best model is based on mean values of lake total-P from three months (TP3/12),
where the last month is month 12 (Dec). The representativity of lake data on total-P is much higher
during late fall and winter than during spring and summer due to the fact that the variability in
phosphorus values is high during spring (peak in water discharge) and summer (peak in production).

114
The same question about time-compatibility could and should be asked for all target variables related
to all model variables in all ecosystems, especially when empirical ELS models are concerned.

Fig. 2.15. The relationship (r2


values) between long-term (3
year) mean Secchi depth (Sec36
in m) and different variants of
lake total-P (in µg/l): totP3/11
denotes a mean value of lake
total-P from three months
where the last month is month
11 (November). Based on data
from 1986 to 1989 from 25
Swedish lakes. From Håkanson
and Peters (1995).

Brief summary:
- All explanatory variables (x-variables) used must be representative in both space and time in relation
to the target variable (y-variable) that is being predicted.
- x-variables should be ranked according to predictive power with respect to the target variable. The
higher the predictive power, the better the model.

2.4. Statistical aspects of regression analysis

2.4.1. The ecometric matrix


It is important to know why a given effect variable, like the Hg-content in fish or the reproduction
damage to key functional fish species, varies among and within lakes. To empirically answer this
question, the actual effect, load and sensitivity variables must be determined for several lakes. The
data must be time and area compatible and these data can then be compiled into an ecometric
matrix (fig. 2.16). In this example, data from 22 different Baltic coastal areas have been used. The
selected target E-variable is a standard effect variable in contexts of eutrophication, the concentration
of chlorophyll-a The load variables are mean concentrations of total nitrogen (TN in mg/m3), inorganic
nitrogen (IN in mg/m3), total phosphorus (TP in mg/m3) and inorganic phosphorus (IP in mg/m3), all
of which could, potentially, influence the primary production and, hence, the values of
chlorophyll-a. Note that the values given in this ecometric matrix are not fluxes (dimension = g
nutrient per time unit) but the resulting concentrations. The data (see table 1.8) were sampled during
July, August and September from several sites in each coastal area (see Wallin et al., 1992 for further
information about these coastal areas, the analysis methods and the sampling program). This is the
period of the year when the primary production is usually the highest. The basic purpose of the
ecometric analysis is to use empirical data like these to make a quantitative ranking of how the load
and sensitivity variables influence the variability in the target effect variables. The task in this case is
to determine how characteristic coastal mean chlorophyll concentrations vary among the given areas.

115
Fig. 2.16. The ecometric matrix in two fashions.
Upper: An ecometric matrix for chlorophyll-a as a standard operational variable for eutrophication effect, load
variables (concentrations of total and inorganic nitrogen and phosphorus) and two morphometric variables
(section area, At; and accumulation area, BA) using data for 22 Baltic coastal areas.
Lower: The general set-up of an ecometric matrix.

So, one important question is:

How much of the variation among the coastal areas in the effect variable (from 0.90 to 4.59 in fig.
2.16) can be statistically explained by the variation in the given load and the sensitivity variables?
This can be tested in several ways, e. g., pair-wise (bi-variate) regression analysis or by stepwise
multiple regression analysis (the latter will be discussed in a subsequent section).

The results from a pairwise regression are given in fig. 2.16. One can note that the r2 value is very high
indeed between chlorophyll and total nitrogen (r2 = 0.89). The r2 values are much lower for the other
tested load variables (0.52 for IN. 0.31 for IP and 0.04 for TP). These r2 values emanate from linear
regressions using the actual values.

116
2.4.2. Confidence intervals and frequency distributions
The regression between values of TN and Chl from Fig. 2.16 is displayed in fig. 2.17A together with
the 95% confidence intervals (CI) for the predicted (individual) y values. Note that these confidence
intervals for the predicted y values (CI = f(SD); SD = standard deviation) are much wider than the
confidence interval for the mean y value (CI=f(SD/√n)).

Fig. 2.17. A. The relationship between actual values for chlorophyll-a (Chl) versus total nitrogen (TN) in 22
Baltic coastal areas. The figure also gives frequency distributions, MV/M50-ratios, 95% confidence intervals for
the predicted y and the mean y, and the r2 value of the regression (0.89).
B. The relationship between log(Chl) and log(TN) illustrating the benefit in of using a log-transformation of
these x- and y-variables. The r2 value is 0.91.
C. The relationship between chlorophyll and inorganic phosphorus (TP) and an illustration that negative values
may appear for the predicted y-variable if the 95% confidence bands are wide apart.
D. The regression equation between log(Chl) and log(TN) and a comparison between modeled values and
empirical data. The figure also gives the "staircase" related to the 95% confidence bands.

From this figure, one can also note that neither the effect variable (Chl) nor the load variable (TN) are
perfectly normally distributed. A good measure of this is the character (the skewness) of the frequency
distribution, and a simple, useful numerical value of the degree of normality of the frequency

117
distribution is the ratio between the mean value (MV) and the median value (M50), which should
be as close to 1 as possible for a normal frequency distribution (disregarding special distributions, like
U-distributions). It is worth noting that skewed distributions yield unreliable r2 values, so a high
degree of normality is indeed desirable.

A large number of water chemical variables are log-normally distributed (Håkanson and Lindström,
1997), and for such variables the log-transformation is a standard procedure to obtain a better
normality of the distribution. From fig. 2.17B, one can note that the MV/M50 ratio is 1.04 for log(Chl)
and 1.00 for log(TN), and that the r2 value has increased from 0.89 to 0.91 by performing this
transformation. This increase in normality and the corresponding increase in the r2 value is one evident
gain. Another major benefit is illustrated in fig. 2.17C (giving the regression between chlorophyll and
inorganic phosphorus), which shows that negative 95% confidence interval values may appear if the
confidence intervals are wide apart, which is the case if the r2 value is low and/or the number of
analyzed data (n) is low. Such statistical relationships will be discussed in the following section, and
this figure is meant to motivate that section.

There is a large number of transformation types available, that can suit different types of frequency
distributions (Box and Cox, 1964). Some common transformations intended to improve the normality
of distributions are the logarithmic (either log10 = log, as in the example above; or ln =logn), or
different exponentials like √x = x0.5, x0.2, x-1, 1/(x+const), etc. The data could also be ranked relative
to the highest and/or lowest numerical value in the series, in which case non-parametric statistical
tests can be performed. Certain transformations (like ex) maximize the weight of high values in
regressions. Others, like log(x), minimize the weight of high values.

Returning to fig. 2.17, the log-log regression between chlorophyll and total nitrogen (log(Chl) =
2.78·log(TN)-6.66) is displayed in fig 2.17D. Very good estimates of chlorophyll-a concentrations
during the summer period in Baltic coastal areas can be obtained from this equation, provided:

1. The coastal areas have been defined according to the topographical bottle-neck procedures
(discussed in section 1.22.) and
2. The TN values fall within the ranges given in table 1.8 (i. e., between 335 and 417 mg TN/m3).
3. The TN values emanate from the same coastal type (glacial, archipelago coasts not influenced by
tides).

This empirical regression model is only applicable under these conditions. All models for complex
ecosystems apply in a defined domain represented by the range and characteristics of the sample used
for deriving an empirical model or used for validating a dynamic model, and the model cannot be used
for other ecosystems without due considerations to this fundamental domain.

Fig. 2.17D also contains a "staircase" related to the 95% confidence intervals for the individual y. The
idea behind this "staircase" (see the coming section) is to show that if modeled values are compared to
empirical data, then the r2 value should be as high as possible and the confidence bands as narrow as
possible.

118
2.4.3. Prairie's "staircase"
Yves Prairie (1996) has produced some very useful results illustrating the practical utility of models
for predictions of individual y values, here referred to as Prairie's: "staircase". If the confidence bands
are wide apart when modeled values are compared to empirical data, then the model can produce
totally useless predictions for individual y values. The usefulness of the predictions is directly related
to the number of steps, or classes obtained in the "staircase" (see Fig. 2.17D). The number of steps is
in turn related to the given r2 value, as illustrated in fig. 2.18.

Fig. 2.18. The relationship


between the number of classes
(NC) and the r2 value when
modeled values are compared to
empirical data, as given by
Prairie's "staircase".

The number of classes (NC) may also be determined by the statistics (the statistical certainty, p, and
the number of data used in the regression, N). If the 95% confidence bands are used, then the
relationship between the number of classes (NC) and the r2 value obtained when empirical data are
compared to modeled values can be approximated by

NC = 1.32 · (1-r2)-0.5 (2.4)

Fig. 2.18 gives the relationship between NC and r2 for different N values, and one can note that if N >
6, eq. 2.4 gives a good description of the relationship.

The important message in fig. 2.18 is that the number of classes increases very rapidly for r2 values
higher than about 0.75, and models yielding r2 values lower than that are more or less useless for
predictions of individual y values (but not necessarily for mean y values in regional modeling
where more uncertain predictions in individual lakes can be accepted, see Håkanson, 1991).

2.4.4. Other statistical concepts and aspects


Regression analyses could be performed for many reasons, e. g., to compare values predicted from
models (generally x) with empirical data (y), to test hypotheses about relationships, and to develop
statistical/empirical models. Many textbooks examine regression analyses (see, e. g., Draper and
Smith, 1966; Cooley and Lohnes 1971; Mosteller and Tukey. 1977; Pfaffenberger and Patterson.

119
1987; Newman, 1993). The aim of this section is to present a very brief summary of some basic
concepts that must be understood in contexts of ELS modeling. The issues of normality and high r2
values have been discussed in the previous sections.

There are linear regressions (y = a · x + b), non-linear regressions (like y = a · log(x) + b) and multiple
regressions (like y = a ·x1 + b · log(x2) + c · eX3 + d). The x- and the y value may either be
either single values or functions (like x = a · z1 + b/z2).

Two fundamental characteristics of any regression analysis are the number of data (N) used in the
regression and the range of the x- and y-variables. They should be considered carefully, because very
different assumptions are implied in building predictive regression models if:

(1) the empirical data base is very small (N < 10), small (N < 30) or large (N > 100), and/or
(2) the range in the x- and y-variable is small relative to the entire possible range of the variable.

The slope of the regression line is given by the constant a in the regression equation y = a·x+b. The
intercept, b, is the y value when x = 0. Regression analysis may be a rather straight-forward and
extremely useful method to find the line that best fits the data. This is often indicated by:

- The best-fit regression line between one x-variable and one y-variable (simple linear regression), or
between several x-variables and one y-variable (e. g., by stepwise multiple regression analysis).
- The correlation coefficient (r), or its squared value, r2. r2 is called the coefficient of determination
or the degree of explanation since perfect fit, 100% statistical explanation, exists when r2 = 1, and no
statistical explanation at all is obtained if r2 = 0.
- The p-level. The probability level gives the statistical significance of the regression. Usually, one
would like to know whether variation in x is significantly linked to variation in y. The r2 value is
highly dependent on the number of analyses (N), the range, the transformation and the residual scatter.
If the other characteristics are constant, the level of statistical significance (p) in the correlation or
regression is related to both N and r2. The statistical derivation of this is left to statistical textbooks.

A nomogram (fig. 2.19) relating the values of r2, N and p, illustrates these relationships. If the x- and
y-variables are normally distributed, then the p value can be read directly from this graph. In this
nomogram, it has been emphasized that the 95% confidence level (p = 0.05) is often used as a default
criteria for significance, but the 90% and the 99% levels are also common.

- The variance and standard deviation of the sample. The variance is simply the average of the
square of the deviations from the mean value; if the mean value of the population is unknown. The
standard deviation (SD) is the square root of the variance; it may be considered as the "average error"
because it is the average distance between the values of y predicted by the regression line and the
observed values of y in the data set used to develop the regression. This statistical concept is
fundamental to defining the confidence limits around predicted individual y values (fig. 2.17) or to
approximating them (since ±2·SD gives the 95% confidence interval for individual values in the
frequency distribution). The standard deviation and variance are relatively independent of range and n,
once the sample size is above about 30.

120
Fig. 2.19. The relationship
between the coefficient of
determination (the degree of
statistical explanation, r2), the
number of data-pairs in the
regression (N) and the statistical
level of significance (p).The
nomogram has been derived by
means of test series and the p-
level of 0.05 is marked since this
is a general (95%) level of
significance. From Håkanson and
Peters (1995).

Another important aspect of regressions is illustrated in fig. 2.20 using the relationship between
chlorophyll and total-N for Baltic coastal areas. The regression equations are NOT the same if the
axes are reversed. The slope is 2.78 in fig. 2.20A (log(Chl) = 2.78·log(TN)-6.66) but 3.03 in fig.
2.20B when log(Chl) is put on the x-axis and log(TN) on the y-axis. This is one crucial aspect of
regressions. A consequence of this, one should always, as common practice, put empirical data on
the y-axis, and modeled values on the x-axis. Likewise, target variables should always be on the
y-axis when plotted against explanatory variables.

Fig. 2.20. Regressions between chlorophyll-a concentrations and total-N (for 22 Baltic coastal areas) where the
x- and y-axes are reversed. Note that there are uncertainties associated with all data on both axes (and not just the
illustrated data point).

121
This information may be crucial in interpretations of empirical data: would the y-variable in fig. 2.20
actually change with a factor of 3.03 or with a factor of 2.78 when the x-variable changes? Using
this operational approach, the answer to that question is 2.78.

It is also indicated in fig. 2.20 that there are always uncertainties related to sampling, transport
analytical procedures and within-area variations for ALL empirical water chemical variables.

Brief summary:
- An ecometric matrix is a practically useful means for structuring information and displaying data.
- A high degree of normality is needed for distributions of all variables in regression analysis.
- Many water quality related variables display log-normal distributions and therefore need to be log-
transformed in a regression analysis.
- To consider a regression reliable, the r2 value should be at least 0.75.
- The statistical significance of a regression depends on the r2 value and the number of samples.
- Target variables in statistical models, and empirical data during model testing, should always be y-
variables in regression plots to get consistent r2 values.

2.5. Variability and uncertainty

2.5.1. Variability within and among aquatic ecosystems


An important objective in ELS modeling is to distinguish between variability in effect, load and
sensitivity variables within and among ecosystems (fig. 2.21). The questions concerning variability
"within and among" is fundamental for understanding issues related to compatible and representative
values of lake variables.

The problem addressed in this section is illustrated in table 2.2 with lake temperature data. The
variability of an ecologically important variable within an ecosystem, CVw, is defined as the
coefficient of variation (= the relative standard deviation, CV = SD/MV) based on data from the given
lake: The CVw of lake temperature is 71.9% for lake 701. The variability among ecosystems (=
lakes). CVa , is defined from the coefficient of variation between data from different lakes: It is
101.29% in the top row of the CVa-column in table 2.2.

Table 2.2. Definition of variability (= coefficient of variation; given in %) within lakes (CVw) and among lakes
(CVa) using data on lake temperature from five lakes.
Lake 701 702 703 704 705 MVa SDa CVa

4.2 8.5 1.5 15.9 0.7 6.16 6.24 101.29


4.0 4.2 0.8 18.8 9.6 7.48 7.07 94.56
6.8 3.1 1.5 16.5 11.8 7.94 6.21 78.24
15.4 1.5 10.0 15.7 14.8 11.48 6.04 52.62
20.1 1.4 11.6 11.0 18.0 12.42 7.32 58.93

MVw 10.1 3.7 5.1 15.6 11.0


SDw 7.3 2.9 5.3 2.8 6.6
CVw 71.9 77.7 103.5 18.2 59.7

Håkanson and Peters (1995) have presented compilations of such CVw- and CVa values for many lake
variables and discussed the rationale for such CV values in contexts of ecosystem modeling. Results
from that compilation are given in fig. 2.22 for six standard lake variables; surface water temperature,

122
pH, total-P, color, Secchi depth and conductivity, based on data from four years of monthly sampling
in 25 Swedish lakes. As expected from this geographically restricted area of the world, there are no
significant differences between CVw and CVa for temperature. The median value for the CV of
temperature is even higher within lakes than among them. In contrast, median values for the CV
values of pH, color, lake total-P, Secchi depth and conductivity (and also Fe- and Ca-concentration,
alkalinity and hardness, see Håkanson and Peters, 1995) are much higher among lakes than within.
The greatest difference in CVa and CVw exists for conductivity; the smallest (except temperature) for
pH and alkalinity.

Fig. 2.21. Illustration of (A) how variable climatic conditions will create time-dependent variations in effects
variables, like 02Sat, and (B) how the variation in mean values in O2Sat among coastal areas may be related to
variations in morphometric characteristics among coastal areas. An aim of this section is to derive an empirical
model yielding an r2 value that is as high as possible when modeled values for 02Sat are compared to empirical
data (C).

If CVa is much greater than CVw, then good predictive models for water chemical variables (like
conductivity and hardness) can be developed from readily available map parameters describing the
catchment area (such as soil type, bedrock type and land use) and the lake morphometry (e. g., form

123
and size parameters), see Håkanson and Peters (1995). The variability within ecosystems is very
important because it determines the representativity of samples, as discussed in the next section.

Fig. 2.22. Box-and-whisker plots (showing 10th. 25th, 50th, 75th and 90th percentiles plus outliers) of
coeficients of variation (CV; in %) for the variability within and among lakes for lake water temperature (°C),
pH, total-P (μg/l), color (Col, mg Pt/l), Secchi depth (m) and conductivity (m S/m). Based on monthly data from
24 Swedish lakes in Håkanson and Peters (1995).

2.5.2. The sampling formula and uncertainties in empirical data


If the variability within an ecosystem is large, many samples must be analyzed to obtain a given level
of certainty in the mean value. There is a general formula, derived from the basic definitions of the
mean value, the standard deviation and the Student's t value, which expresses how many samples are
required (n) in order to establish lake mean value with a specified certainty (Håkanson, 1984b):

n = (t·CV/L)2 + 1 (2.5)

where t = Student's t, which specifies the probability level of the estimated mean (usually 95%;
strictly, this approach is only valid for variables from normal frequency distributions), and CV denotes
the coefficient of variation within a given ecosystem (we will not discuss CVa values in the following,
so CVw is, for simplicity, written as CV hereafter). L is the level of error accepted in the mean value.

For example, L = 0.1 implies 10% error so that the measured mean will be expected to lie within 10%
of the expected mean with the probability assumed in determining t. Since one often determines the
mean value with 95% certainty (p = 0.05), the commonly used t value is 1.96 (from statistical tables),
so that eq. (2.5) can be re-written as:

n = (1.96·CV/L)2 + 1 (2.6)

124
The relationship between n, CV and L is illustrated graphically in fig. 2.23. This figure also gives
typical CV values for many lake variables discussed in this book, like radiocesium concentration in
fish, water and sediments (from Håkanson, 1998). Hg-concentrations in fish, and many water chemical
variables (from Håkanson and Peters. 1995). Many of these CV values will be used in subchapter 2.8
in uncertainty tests using Monte Carlo simulations.

Fig. 2.23. The sampling


formula. Nomogram
showing how many
samples must be analysed
(n) to establish a
characteristic mean value
with a given uncertainty or
error (L) and a given
confidence level (p=0.05).
CV is the coefficient of
variation. Characteristic
CV values relaIed to mean
annual dam for different
variables are also given, L
= 0.1 (i. e., 10% of the
mean).

If the CV is 0.35, about 50 samples are required to establish a lake-typical mean value for the given
variable provided that we accept an error of L = 10%. It one accepts a larger error, e. g., L = 20%,
fewer samples would be required.

Since most variables in most lakes and coastal areas have CV's between 0.1 and 0.5 (fig. 2.23 and
table 2.3), one can calculate the error in a typical estimate. If n = 5 and CV = 0.33, then L is about
33%. Since few monitoring programs take more samples, this calculation has profound implications
about the quality of our knowledge of aquatic systems. It shows that for most water variables, existing
empirical estimates are only rough measures of the area-typical mean value. This is especially so for
total-P (CV = 0.35), and to a lesser extent for Secchi depth (CV = 0.15), conductivity (CV = 0.1) and
pH (CV = 0.05).

The same argument can be made for any given variable for any given ecosystem. Table 2.3 gives
typical CV values for some important variables in coastal ELS modeling. The CV values are often on
the same order of magnitude for lake and coastal area variables, e. g., Secchi depth, CV ≈ 0.15-0.2 and
chlorophyll, CV ≈ 0.25. Note, however, that CV for total-P is significantly greater for lakes (CV ≈
0.35) than for coastal areas (CV ≈ 0.16). The main reason for this may be that total-P is the short-term
limiting nutrient in lakes where it participates in many biological reactions. This increases the
variability.

One reason for some of the high CV values has to do with the fact that there are large differences in
analytical reliability for different variables (see Håkanson et al., 1990b. for further information).

125
The average uncertainty related to analytical procedures (CV) is only about 0.075 for the Hg-
determination. Lake pH can also be determined with great reliability. The average analytical CV value
for pH is about 0.02. This represents the combined effects of errors in all phases of sampling, sample
preparation and analysis. Determinations of conductivity, hardness (CaMg) and the Ca-concentration
are also generally highly reliable. Color, Fe concentration, total-P concentration and alkalinity have
much higher methodo1ogical/analytical CVs (0.15-0.2).

Variable n CV L (%)
Secchi 12 0.19 11.4
Chl 7 0.25 20.3
TN 7 0.13 10.3
IN 7 0.31 25.0
TP 7 0.16 13.0
Table 2.3. Average number of samples in individual IP 7 0.28 22.6
coastal areas (n), coefficients of variation (CV =
SedS 5 0.58 46.1
SD/MV) and the calculated statistical error (L) at
SedB 5 0.50 39.8
95% confidence level for different water quality
and sediment hap variables (based on Wallin et al., O2-conc. 12 0.26 15.1
1992). O2Sat 12 0.25 14.8

It is important to remember the uncertainty in empirical data when one derives empirical models
or calibrates and validates dynamic models using empirical data. Since one rarely has very reliable
empirical data, one cannot expect to obtain models which explain all the variability in the target
effect variable. The uncertainty in the empirical determinations of variables for which relatively few
samples have been analyzed may produce marked divergences between modeled and empirical data.
In such cases, wide divergences may not necessarily indicate errors and deficiencies in the models but
could reflect deficiencies in the empirical base. This is the focus of the next section.

Brief summary:
- Variables with higher variability among lakes than within lakes can be predicted with high certainty
using map parameters.
- The mean value error depends on the number of samples taken and on the coefficient of variation for
the variable in question.
- Empirical data are not "cut in stone" but their uncertainty and variability could and should be
accounted for.
- Many factors (sampling and analysis methods, physical, biological and chemical processes, etc.)
influence empirical data.
- Model uncertainty should be compared with the uncertainty in empirical data.

2.6. Principles determining the predictive success of ecosystem models

If an ecosystem model is tested against an independent set of data, the achieved r2 value when modeled
values are compared to empirical data (y) will depend on the uncertainty in the empirical y value (i. e.,
the uncertainty in the y-direction) and on the structuring of the model, i. e., which processes and model
variables are accounted for, how this is done, as well as the empirical uncertainty of the model
variables (i. e. the uncertainty in the x-direction).

126
The aim of this section is to highlight fundamental principles and factors regulating the predictive
success of ELS models. Three types of r2 values will be discussed:

1. The highest reference r2 (rr2), which will be defined here.


2. The empirically based highest r2 (re2), which is determined from a regression analysis when two
parallel empirical data sets of y values are compared (Empl vs Emp2).
3. The highest achieved r2 when modeled y values are regressed against empirical y values (r2).

Ecosystem models can never be expected to yield high predictive power if the target variable (y)
cannot be empirically determined well and/or if the model is based on driving variables (x)
yielding a high coefficient of variation (CV).

A key question in predictive ecosystem modeling is: How high is the highest potential predictive
power for a given target variable y? It is evident that many factors have to be considered, e. g.,
sampling (such as the number of samples), analysis (e. g., the precision in determining y), model
structure (e. g., how and which x-variables are included), the reliability of the model variables and the
statistical methods used to define predictive success.

In this section, these issues will be discussed from a top-down approach and data from lake
ecosystems will be used to exemplify the principles. Top-down here means that we will start with the
target variable (the y-variable) and try to rank the factors influencing the predictive success of y.
We will also use the r2 value (from regressions where modeled values are compared to empirical data)
as a standard criterion of predictive success since this is a widely used concept in ecosystem modeling.
The benefits and disadvantages with the r2 value are probably better known than for, e. g., predictive
power, functional distance and/or CI.

2.6.1. The highest possible r2 from Emp1-Emp2; re2


One way to determine the highest possible r2 of a predictive model is to compare two empirical
samples. The variables in these two samples should be as time and area compatible as possible: they
should be sampled, transported, stored and analyzed in the same manner.

Fig. 2.24 covers some fundamental concepts in this context. Fig. 2.24A gives empirical data for the
target variable y on both axes, i. e., Empl vs. Emp2, when the dataset has been divided into two
randomized subsamples that capture the same uncertainty as the full dataset does. There are
uncertainties for all these values, and this case concerns a uniform CV value of 0.35. The
uncertainty associated with the given target variable is illustrated by the uncertainty bars, which are
equally large in all directions because Emp1 and Emp2 describe the same thing. This uncertainty will
evidently influence the result of the regression, such as the r2 value and the confidence intervals. If CV
for y is large, one can NOT expect a model to predict y well.

The r2 value from a regression between Emp1 and Emp2 as in this example (fig. 2.24.A) is the
definition of re2, "the empirically highest r2".

Fig. 2.24B shows a normal model validation when modeled values are put on the x-axis. The
empirical uncertainty associated with y remains the same on the y-axis but the uncertainty in the x-
direction is related to the uncertainty associated with the model structure and the uncertainty of the

127
model variables (x). Generally, one would expect the model uncertainty to be larger, or much
larger, than the uncertainty in empirical data (on the y-axis in Fig. 2.24B).

This means that the r2 value in the regression in fig. 2.24B is likely to be lower than re2 (the r2 value
obtained in the Empl-Emp2 comparison in fig. 2.24A). The residual value, R2 = 1-r2, must then be
higher in fig. 2.24.B than in fig. 2.24A. re2 may be seen as one approach to answer to the crucial
question of how far it is possible to reduce residual uncertainty. The minimum residual uncertainty
in modeled data can only be achieved by building/structuring the model in the best possible manner by
accounting for the most important processes/model variables and by omitting the relatively
unimportant processes/variables which add more to the model uncertainty than to the predictive
success (see Håkanson and Peters, 1995; Håkanson, 1995c for a more detailed discussion on optimal
model size).

Fig. 2.24. Illustration of some


fundamental concepts related to the
question of "the highest possible r2"
of ecosystem models.
A. Empirical data for the target
variable y on both axes, i. e., Empl
vs. Emp2.
B. Empirical data on the y-axis and
modeled values on the x-axis.

Fig. 2.24 illustrates a rather simple scenario for empirical/statistical regression models which produce
one y value for one ecosystem (e. g., a lake). However, many ELS models yield time-dependent
predictions (time series) of y, where the data in the time series are not independent of each other.
Fig. 2.25 illustrates this situation. We have empirical data of the target y-variable and the empirical

128
uncertainties associated with y (the CV value is also in this case set to 0.35) and the series of model-
predicted data (which can be expected to be even more uncertain) on the y-axis and time on the x-axis.

Fig. 2.25. A comparison between one empirical data series with uncertainty intervals (CV = 0.35) and modeled
values in a time series from a dynamic model.

In fig. 2.25, we have illustrated a situation where the predictions fall within the empirical confidence
bands for all months except for June. July and August, i. e., a situation where the model structure is
systematically inadequate. By changing (excluding and/or including) equations and/or model
variables, it may be possible to obtain a better fit between empirical data and modeled values for the
summer period. The fit can be expressed by the r2 value, but also by various measures of the difference
between modeled values and empirical data.

Such a comparison is given in fig. 2.26A. We can note that the r2 value is very low in this example
(0.012) in spite of the fact that the model gives rather accurate predictions - illustrated by the fact
that the mean error (Diff = (M-E)/E, M = modeled value, E = empirical data) is just 0.11, i. e., 11%.
This illustrates a well-known drawback with the use of r2; r2 depends on the number of data and the
range of the data in the regression. The r2 value increases with a widening data range if the
relative prediction error remains the same.

This is illustrated in fig. 2.26B. The data from fig. 2.26A is given by the cluster called "Lake 3". In
contexts of ecosystem modeling, i. e., when the ultimate aim is to have a general model covering the
entire range of the target variable y, it is evident from fig. 2.26 that the r2 value is not an adequate
statistical indicator for time series of dependent data from a certain site with a small y-range. In
such cases, one can preferably use, e. g., the Diff value. The r2 value could, on the other hand, be used
as a criterion for the fit for the entire data set. Fig. 2.26B illustrates that this hypothetical model
yields very good predictions (r2 = 0.979) over the entire range.

To address the key question about the highest predictive power and the relative role of the empirical
uncertainties in y and the model structure, fig. 2.26C highlights an important aspect. A time series
from a fourth lake ecosystem (called Lake 6) has been included along with the three lakes given in fig.

129
2.26B. One can note that the model structure of this dynamic model cannot cope with the conditions
prevailing in Lake 6. The r2 has also dropped to 0.645. A comparison between two data sets for all
four lakes for y (Empl vs Emp2, such as in fig. 2.24.A) gave an r2 value of 0.92 under defined
statistical conditions, like CV for y = 0.35. This means to that it would be possible to improve the
model significantly, from r2 = 0.645 to an r2 higher than 0.9 if a better model structuring would be
used.

Fig. 2.26. A. A regression between empirical data (Emp3


on the y-axis) and modeled values (Mod3 on the x-axis).
The r2 value for these 12 data is 0.012 and the Diff value
0.11.
B. A comparison between data series including two
additional lakes (Lakes 4 and 5). The overall r2 value has
now increased to 0.979.
C. The same three lakes and a new data series from Lake
6. The model gives poor predictions for Lake 6 but good
predictions for the other three lakes, so the model structure
is at least partially falsified and should be improved. A
direct comparison between two empirical data series (Empl
versus Emp2) yields an r2 value (re2) of 0.92, so the poor
results for Lake 6 is not due to deficient empirical data but
to a suboptimal model structure.

To illustrate the basic approach to determine re2, we will use fig. 2.27, based on data from 70 Swedish
lakes (from Håkanson et al., 1990a). Data concerns the target operational variable in mercury research,
the Hg-concentration in fish for human consumption (1-kg pike; values in mg Hg/kg muscle;
abbreviated as Hgpi). At least 4 fish are included in each sample, i. e., at least 8 fish per lake. The
Emp1-Emp2 regression in fig. 2.27 gives at hand that re2 for Hgpi in this range of lakes is 0.86.

The following sequence (compiled from Håkanson and Peters, 1995) of values for re2 regarding some
water quality related variables has been determined from two sets of mean values, each representing 6
samples from different months from 1986 from 25 lakes.

Variable Temp Total-P Secchi Fe Ca & pH Alk, color & CaMg


re2 0.76 0.85 0.90 0.95 0.96 0.99

130
Temperature data yielded, as might be expected, the lowest value (0.76), and color one of the highest
(0.99). These data also suggest something about the reactivity and/or temporal variability of these
variables, which is important information in establishing representative values for larger areas and
longer periods of time (like annual or monthly mean values). The more reactive and changeable the
variable, the more difficult it is to establish representative empirical data of the given variable.

Fig. 2.27. Determination of "the empirically


highest r2" (re2) for the Hg-concentration in 1-
kg pike (Hgpi) from a regression between two
parallel samples from 70 Swedish lakes.

The main point here is that one should never hope to explain all of the variation in any ecological
variable. It is interesting to note that total-P, which is a fundamental state variable in practically all
lake contexts, displays rather high variability and low re2 (0.85).

2.6.2. Highest reference r2, rr2


The highest reference r2, rr2, presented here is meant as a simple, practical tool in ELS modeling to
obtain a highest reference r2 only to related to the variability in the target y-variable. rr2 is thus
complementary to re2 and is used for the same purpose: to estimate how high predictive power we
can expect to get for a model which targets a certain water variable.

The task is to define a formula of the type:

rr2 = f(CV) (2.7)

where CV is the coefficient of variation for the y variable.

The derivation of the following expression for rr2 (from Håkanson and Peters, 1995) is based on
sampling formula (eq. 2.6) and the equation for the 95% confidence interval (CI) for individual y
values from a number (N) of independent validations. CI is a function of N, r2, and the range of y,
while the sampling formula provides the mean value error level (L) as a function of CV and n. If CI is
set equal to L and if n and N are set to 10, and if the minimum value in the range (ymax-ymin) is
assumed to be small in comparison to ymax, and if ymax is set to 1 (which is valid for relative values),
then:

131
rr2 = 1 - 0.66 · CV2 (2.8)

This is the definition of the requested highest reference r2 value, rr2. The equation is graphically
shown in fig. 2.28. It is valid for actual (i. e., non-transformed) y values.

Fig. 2.28. The relationship between


"the highest reference r2" (rr2) and the
characteristic coefficient of variation
for variability within (CV)
ecosystems.

It is evident that there are many assumptions involved in this derivation. The whole idea is to try to
make simplifications motivated from the perspective of ELS modeling where it would be very useful
to have one reference r2 related to the inherent uncertainty in the target y-variable. The practical use of
rr2 and re2 will be shown in the following parts of the book, starting with the next section.

2.6.3. Comparing model predictions with re2 and rr2


The OECD model (see fig. 1.64) is an empirical modification of the basic Vollenweider model, and it
has yielded the hitherto highest r2 value of 0.86 (for log-transformed data) when tested for 87 lakes
covering a wide range (TP concentrations from 2.5 to 100 μg/l = mg/m3) for lakes with theoretical
water retention times from 0.1 to 100 years). The r2 value for actual (untransformed) data is probably
lower, about 0.82-0.83. The two driving variables in the model are the mean annual TP concentration
of tributary water (Cin) and the water retention time (Tw). This is a very simple model indeed, and, in
fact, too simplistic for many important lake eutrophication problems (for which this model is
nevertheless often used).

Fig. 2.29 illustrates the determination of re2 for lake TP concentrations. Two parallel sets of empirical
annual mean values are compared. One series (Emp1) of data from months 1, 3, 5, ..., the other data
series (Emp2) of data from months 2, 4, 6, ... Thus, there are 6 data (n = 6) for each mean value in
Emp1 and Emp2 and 25 (N = 25) lakes in the regression.

The empirically based r2 value, re2, is 0.85. It is evident that this value depends on the available data
set (on n and N). This value can be compared to the r2 value obtained for the actual data (and not the
log-uansformed data) for the OECD model of about 0.82-0.83. A quick comparison from a statistical
perspective would then indicate that the OECD model is "very good". There are, however, as already
stressed, several ecosystem arguments against simplistic empirical models, such as the OECD model.
This will be further elaborated using uncertainty analysis in section 2.7.

132
Fig. 2.29. Illustration of "the empirically
highest r2" (re2) for the total phosphorus
concentration. This regression is based on
parallel data sets of 6 samples (n) from 25
Swedish lakes (N). Revised from Håkanson and
Peters (1995).

As emphasized before, the highest empirically based r2 value of 0.85 depends very much on the
quality of the two empirical data sets (Emp1 and Emp2).

The highest reference r2 (from eq. 2.5), is rr2 = 1 - 0.66 · 0.352 = 0.93 (if the characteristic CV for TP is
set to 0.35; see fig. 2.23), and this may be a more relevant general reference value for the highest r2
value one can expect to achieve for models to predict TP concentration in lakes. Table 2.4 gives a
compilation of r2, rr2 and re2 values for effect variables and ELS models discussed in this book and also
question marks where data are missing.

Table 2.4. Compilation of highest possible "empirically based r2 values" (re2), "highest reference r2 values" (rr2)
and highest achieved r2 values (r2) from models (empirical and/or dynamic) discussed in the PER-analysis for the
given chemical threats to aquatic ecosystems.
Chemical Lake (L) Effect variable CV rr2 re2 Highest Static (s) or
threat or coast mo- dynamic (d)
(C) deled r2 model
Acidification L Change in roach biomass (?) ? ? ? ? ?
Mean annual pH 0.05 0.998 0.96a 0.42ab s
Mercury L Hg conc in fish 0.25 0.96 0.86cd 0.85d s, d
Radiocesium L Cs conc in fish 0.22 0.97 ?d 0.98d d
Eutrophication L Chl-a conc 0.25 0.96 ? 0.96e s
Eutrophication C Chl-a conc 0.25 0.96 ? 0.91c s
Eutrophication C O2 saturation 0.25 0.96 ? 0.93c s
Metals (Cd, ? ? ? ? ?
Pb, etc.)
a) Håkanson and Peters (1995); b) based on map parameters; c) this book; d) Håkanson (1999); e) Peters (1986).

Brief summary:
- Two quantitative concepts, re2 and rr2, have been presented that can be used to determine the highest
potential r2 values for regressions between modeled and empirical data.
- re2 is based on a comparison between empirical datasets, Emp1-Emp2.
- rr2 is based on the CV value.
- Both re2 and rr2 differ for different water quality variables, which means that these can be modeled
with a varying degree of certainty.

2.7. Dynamic and static ecosystem modeling

2.7.1. The classical ELS model - lake eutrophication


This section deals with the classical modeling approach for ELS models using the basic mass-balance
model and an empirical regression for the target effect variable. The following section will give the
basic elements of empirical ELS modeling. The concepts discussed in these two sections are

133
fundamental for a proper understanding of how ELS models are built, their structure, potential and
limitations.

The basic mass-balance model for the flow of matter, or in this example total phosphorus (TP), to and
from a lake, may be described by the following differential equation (see also fig. 2.30):

V · dC/dt = Q · (Cin-C) - KT · V · C (2.9)

where
V = the lake volume (in m3);
dC/dt = the change in substance concentration (dC; in g/m3 or similar) in lake water per unit of time
(dt; usually in weeks, months or years);
C = the concentration (in g/m3 or similar) of the substance in lake water; often, as here, set equal to the
outflow concentration,
Cin = the average concentration (in g/m3 or similar) of the substance in tributaries (flow adjusted so
that all tributaries are given weight to the average according to their water flux contribution)
Q = the tributary water discharge to the lake (m3 per time unit)
KT = the turnover rate of the given substance C in the lake; this rate has, as all rates, the inverse of
time (1/week, 1/month, 1/year or similar) as a dimension

Fig. 2.30. The basic components of the


mass-balance equation of a substance for
a lake. Q is the water discharge to the
lake (Qin = Qout). The concentration of
the substance in the inflow is abbreviated
Cin and in the lake and its outflow C. KT
is the rate of sedimentation (1/time). v is
the settling velocity (length unit/time), V
is lake volume, area is lake area and dC
the change in concentration during the
time dt. From Håkanson and Peters
(1995).

The lake may be envisioned as a reactor tank (fig. 1.65) with complete mixing during the calculation
time (dt). The simplest way of solving this equation is to assume steady-state conditions. That means
that one conserves mass and puts Qin = Qout = Q.

The lake water retention time (or residence or turnover time, Tw, in time units; generally days,
weeks, months or years) is defined as the ratio between the lake volume and the water discharge, i.
e.: Tw = V/Q. This is the time it takes to fill a compartment of volume V if the water discharge to
the compartment is given by Q; alternatively, Tw can be seen as the average time that a water
molecule remains in the lake. Q could be taken from time-series of measured values or from models
(see, e. g., Håkanson and Peters, 1995; Abrahamsson and Håkanson, 1998). If mean (monthly, annual,
etc.) values are used for Q, one generally refers to Tw as the theoretical water retention time.

134
The retention time of a chemical or a given suspended particle (Tr; in time units) is defined in the
same way, as:

Tr = V · C / (Q · Cin) (2.10)

The relationship between the water retention time (Tw) and the substance retention time (Tr) is
important. Tw is by definition equal to Tr for water and for conservative or non-reactive substances,
i. e., substances which do not change (settle, evaporate or transform in the lake). Tr < Tw for most
allochthonous (=coming from the catchment) particles and for pollutants which are transported to the
lake from the catchment and then distributed in a typical pattern with lobes of decreasing
concentration gradients with distance from the source of pollution, or from the tributary river
mouths.

By making a steady-state assumption, i. e., by setting dC/dt = 0, one may solve eq. 2.9:

C = Q · Cin / (Q + KT · V) (2.11)
or C = Cin / (1 + KT · Tw) (2.12)

Many management models for TP do not account for internal loading, i. e., the transport of TP from
sediments back to water, since this is governed by so many factors that have historically been
considered difficult to quantify. Eq. 2.12 may seem like a straight and simple approach to estimate C
(the lake concentration of TP) but this approach would generally give poor predictions because:

1. It requires data on the net settling rate for total phosphorus (KT, or Rsed) which is very difficult to
determine, since, by definition, only the particulate phase of phosphorus, and NOT the dissolved
phase, is physically involved in the sedimentation process, and KT for particulate phosphorus is not
constant but variable.

2. It requires data on the tributary concentrations of TP ( Cin = CTPin), which are less difficult but costly
to get since they also vary considerably, and a large number of important but small tributaries and
groundwater inflow makes the task even more tedious. To obtain reliable, representative monthly or
yearly mean values for a lake model, a large number of inflow samples must be collected and
analyzed.

In one of the first nutrient load models for lakes, presented by Vollenweider (1976), the basic mass-
balance equation was somewhat altered. Empirical data indicated that a better prediction of C for TP
could be obtained by the following formula:

CTP = CTPin / (1 + √Tw) (2.13)

In this expression, there is no KT, and instead of using KT · Tw, one has √Tw = Tw0.5, which gave
better predictions when empirically tested. Instead of accounting for more processes, Vollenweider did
it the other way around: He simplified the basic mass-balance approach, omitted KT - or, rather, he
approximated KT with 1/√Tw .

135
In eq. 2.13, one would still need data on CTPin One way to circumvent the demanding requirement of
having CTPin as a driving parameter is to use the following expression:

CTPin = SRP · TW / Dm (2.14)

That is, CTPin is assumed to be a function of SRP, the specific runoff of TP from land per unit of area
and unit of time (generally mg TP/(m2·yr), multiplied by Tw and divided by the mean depth (Dm) to
get the proper dimension for CTPin (in mg TP/yr). Equations 2.13 and 2.14 may occasionally provide
useful predictions of CTP, but still, one would need data on SRP, which in fact may be just as difficult
to obtain as it is to get data on CTPin.

A better prediction of CTPin can in some cases (but not systematically) be obtained by the following
model (OECD, 1982: see fig. 1.64):

CTP = 1.55 · (CTPin / (1 + √Tw))0.82 (2.15)

It is run by means of the same driving parameters as the Vollenweider model but uses three
empirically based constants instead of one; 1.55, 0.5 and 0.82. CTPin and CTP represent mean annual
concentrations.

As already mentioned, there are several major drawbacks with models of this type:
1. They use mean annual values, hence they do not account for seasonal variations in important
fluxes, rates and state variables.
2. They do not account for internal loading, which is very important, particularly in recovering or
degrading lakes where conditions are very far from steady state; i. e., the models predict poorly in
many lakes for which the load models are often needed the most
3. They do not differentiate between bioavailable (= dissolved) and non-bioavailable (=particulate)
fractions of phosphorus. Operationally, this separation is generally done by filtering through a 0.45 μm
filter.
4. They treat the important question concerning the lake water retention rate in a very simplistic
manner, and
5. They do not provide a direct quantitative link to ecological effect variables.

Table 2.5 lists 17 models of the Vollenweider-type for predicting mean annual lake TP from different
empirical variables and morphological parameters. The reason for this multitude of models is that
these models are, evidently, too crude and uncertain to be useful for many lake management
applications in individual lakes, like evaluations for permits to run fish farms. Many of the models in
table 2.5 have been developed for certain regions or for specific lake types in order to increase model
precision. Despite the many attempts to produce better models, only minor improvements have been
achieved compared to the original Vollenweider model. It therefore seems unlikely that this kind of
model approach can ever be significantly improved. A more appropriate model structure is needed!

TP is a crucial variable in lake management because many other target variables can be predicted from
it. The maximum volume of phytoplankton (PP in mm3/l) can be calculated from CTP from the
following regression (see fig. 1.66):

136
PP = 0.01191 · CTP1.512 (2.16)

The two equations 2.15 and 2.16 constitute one variant of a classical ELS model for lake
eutrophication based on both dynamic modeling and empirical regression analysis. The critical PP
value of 5 mm3/l and/or the alarm level of PP = 10 mm3/l may be used as general Ecrit value in the
PER-analysis, although the best value would be obtained by comparing present PP values with pre-
industrial values at each site, since natural ecosystems may also be naturally eutrophic.

Table 2.5. Vollenweider-type models to predict annual mean lake TP concentrations (mg/m3). L = areal TP load
(mg/m2·yr). Dm = mean depth, KT = lake water retention (or sedimentation) rate (dimension 1/yr). KP = TP
retention (or sedimentation) rate (1/yr), q = areal water load (m3/(m2·yr)), v= sedimentation velocity (m/yr), a =
areal calculations and vl = volumetric calculations. From Meeuwig and Peters (1996).
Model Calculations based on
TP1=0.8·L/Dm·(0.0942·(L/Dm)0.422+KT -
TP2=0.49·L/(Dm·(0.0926·(L/Dm)0.510+KT -
TP3=L·(1-(v/(v+q)))/z·KT where v=11.26+1.2q a
TP4=L·(1-(v/(v+q)))/z·KT where v=12.4 a
TP5=L·(1-(KP·(KP+KT))/z·KT where KP=0.94 vl
TP6=L·(1-(KP·(KP+KT))/z·KT where KP=0.162·(L/Dm)0.458 vl
TP7=L·(1-(KP·(KP+KT))/z·KT where KP=0.129·(L/Dm)0.549 vl
TP8=L·(1-(KP·(KP+KT))/z·KT where KP=10/Dm vl
TP9=L·(1-(v/(v+q)))/z·KT where v=2.99+1.7q a
TP10=L·(1-R)/z·KT where R=1/(1+KT0.5) vl
TP11=L·(1-R)/z·KT where R=1/(1+0.747·KT0.507) vl
TP12=L·(1-R)/z·KT where R=0.426·e(-0.271·q) +0.574·e(-0.00949·q) a
TP13=L·(1-24/(30+q))/z·KT vl
TP14=L·(1-R)/z·KT where R=0.201·e(-0.0426·q) +0.574·e(-0.00949·q) a
TP15=L·(1-R)/z·KT where R=1/(1+0.614 · KT0.491) vl
TP16=0.603·L/(Dm·(0.257+KT)) -
TP17=L·(1-KP/(KP+KT))/z·KT where KP=0.65 vl

The areal and temporal extent of the selected Ecrit values may then be calculated, e. g., using a map of
Sweden with characteristic CTP values for many lakes from the national monitoring program, and
using the models just given together with a modern GIS-program like ArcGIS (GIS = Geographical
Information System). The user must then also define which PP values may be set to E = 9 (very large
ecosystem effects) and to E = 10 (total collapse of a natural ecosystem).

2.7.2. ELS modeling of coastal eutrophication


This section contains an example of how the cause and prevention of an environmental threat, coastal
eutrophication, can be quantitatively predicted by means of ELS modeling. The target variable is
O2Sat, the oxygen saturation.

It should be stressed once more that it is NOT possible to make simple adjustments of lake models of
the Vollenweider- and OECD-types to coastal marine areas, because the physical conditions are
fundamentally different there compared to lakes (see 2.2). The surrounding sea often has a profound
influence on coastal ecosystems. A typical water retention time for lakes is about 1 year; for coastal
areas it is about 2-4 days! A low mean depth in a lake implies significant resuspension and a high
internal loading; in an open coast, it implies a low internal loading of nutrients since the bottoms
would be dominated by coarse and hard deposits with a low nutrient content (see fig. 1.6).

137
From the complex hydrodynamical and sedimentological conditions described in 2.2.2., one might get
the impression that the complexity prevents models of high predictive power to be developed. This
will be tested in the following section. The working hypothesis is that many factors can, potentially,
influence the oxygen saturation of the deep water (O2Sat). By using empirical data and a structured
analytical procedure, the aim is to show that one can make a ranking of the factors influencing the
variability among coastal areas in characteristic mean O2Sat. This is also a general structure which
can be used to derive any empirical ELS model.

The utilized data are summarized in table 1.8. Fig. 2.31 illustrates the organization of the project that
generated the data used in this work, starting with definitions of coastal ecosystem boundaries at the
topographical bottlenecks, via digitatisation of bathymetric charts, GIS calculations of morphometric
data, derivations of empirical models for key parameters like theoretical surface water retention time
(Ts), deep water retention time (Td) and bottom dynamic conditions to the development of ELS
models.

The mean summer oxygen saturation in the bottom water (i. e. the water beneath the thermocline) will
be used as the target variable and the operational effect variable (O2Sat in %). From fig. 2.32 one can
see that there exists a very close and logical relationship between the O2 concentration (in mg O/l) and
the O2 saturation (in %). This means that the simple approximation O2-conc = 0.1·O2Sat may be used
to predict reasonable values for the O2 concentration.

When the O2 concentration is lower than about 2 mg/l, and O2Sat lower than about 20%, many key
functional benthic groups die or leave (fig. 1.61). It is then, naturally, convenient to develop ELS
models based on an operational chemical effect variables like O2Sat but then one must demonstrate the
biological/ecological significance and relevance of such variables. This is quite clear from fig. 1.60.

The general avenue (see Håkanson and Peters, 1995) to derive an empirical ELS model will be
followed in this section. There exist data on the selected y-variable, O2Sat, from 23 Swedish Baltic
coastal areas. The following model derivation is based on these arguments:

1. One important aspect of statistical analyses of empirical data like these is that they enable a
quantitative ranking of the factors that actually influence the variability of the target y-variable. All
factors cannot be of equal importance, and this exercise will provide such a ranking. Results will,
however, only apply for areas belonging to the given coast type (glacial coasts without tide) and NOT
for all coast types around the world.

2. From the already cited literature (see section 1.6), one would expect that nitrogen would be at least
as important as phosphorus for eutrophication effects in Baltic coastal areas. Is this relevant also for
O2Sat?

3. It is possible to derive practically useful empirical models for O2Sat. The arguments given in section
2.4.3 (see fig. 2.18) clearly demonstrate that models yielding r2 values lower than about 0.75 (and p
values higher than about 0.05) are practically useless for predictions in individual coastal areas.

138
Fig. 2.31. Illustration of
the structuring of coastal
ELS modeling (modified
from Wallin et al., 1992).

4. It is evident that variables regulated by climatic conditions, like long-time and short-time variations
in temperature, precipitation, deposition of nitrogen and resuspension, would be important for the

139
variability in O2Sat within coastal areas (see fig. 2.21). But what relative importance should one
give to the constant morphometric characteristics, not in relation to site-specific conditions, but to
the mean, area-characteristic conditions during one month and/or one season of the year?

Fig. 2.32. The relationship between


empirical mean values of the O2
concentration (in mg O2/l) and O2
saturation (in %) from 23 Baltic
coastal areas for data from the summer
period (July-September).

If it were possible to quantify this, then it would also be possible to say that x % of the variability in
mean O2Sat could be attributed to the defined, constant morphometric parameters and the rest of
the variability to variables. Such knowledge, and the possibility to predict and quantify, would
certainly provide a better basis for understanding causation and possible remedial strategies.

One very important prerequisite for this model concerns the definition of the coastal ecosystem, i. e.,
where to place the boundaries toward the sea and/or adjacent coastal areas. It is most important
to use a technique that provides an unambiguous, ecologically meaningful and practically useful
definition of the coastal ecosystem boundaries. The coastal limitation lines are, in this context
operationally defined at the topographical bottle-necks so that the exposure (Ex) toward the sea and
adjacent coastal areas is minimized (see section 1.2.1 and particularly fig. 1.6).

The aim here is to predict mean summer values of O2Sat for entire and well-defined coastal areas. It
should be noted that summer is the time of the year when the climatic conditions are likely to create
the lowest values for O2Sat in this part of the world, so it is interesting to try to predict the worst
possible conditions.

Procedure checklist:
1. Frequency distributions. O2Sat appears with a negatively skewed frequency distribution (fig.
2.33A). This means that suitable transformations to obtain a more normal distributions is, e. g.,
log(O2Sat+0), since O2Sat may be equal to 0 and log(0) is not mathematically defined, or arcsin(O2Sat
/100), see fig. 2.33B. But other transformations, like (O2Sat)2, may also produce a more normal
frequency distribution than the original one given in fig. 2.33A. The degree of normality is here
given by the ratio between the mean value (MV) and the median value (M50). This value should
be as close as possible to unity. It is 0.94 for the untransformed values of O2Sat and 1.02 for the
arcsin-transformed values. During periods of very high primary production, O2Sat may attain values

140
higher than 100% (up to about 125%). If the arcsin transfomation is used, O2Sat values higher than
100 should be set to 100.

Fig. 2.33. Histograms for O2Sat (figures A and B), total-P (fig. C), total-N (fig. D), a load function,
log(TN+10·TP) (fig. E), the mean depth [log(Dm); fig. f], modeled values for the theoretical deep water retention
time [log(1+Tdmod); fig. G], and gross sedimentation in near-bottom sediment traps [log(SedB); fig. H]. Based on
data from 23 Baltic coastal areas. The ratio MV/M50 indicates the degree of normality of the frequency
distribution.

The model is meant to be used for predictions for the entire population of coastal areas out of which
these 23 areas have been selected as a sample. This means that the transformations that are likely to
best represent the population should be used in this predictive model.

From fig. 2.33, one can also see the frequency distributions of some interesting load variables:

- total-P (TP, fig. 2.33C; MV/M50 = 0.991),


- total-N (TN, fig. 2.33C; MV/M50 = 0.995),

141
- a load function [log(TN+10·TP), fig. 2.33E, MV/M50 = 0.998]. Other load functions, like
(TN+x·TP), (TN·TP)x have also been tested, but these tests are not given here. The figure gives
frequency distributions also for some other variables which will be used later:
- the mean depth Dm in m; MV/M50 = 0.97 for [log(Dm)], which is a well known and useful
morphometric form parameter,
- modeled values for the theoretical deep water retention time (Tdmod in days; fig. 2.33G. MV/M5O
= 1.53). This model emanates from Persson and Håkanson (1996). It is a regression model based on
extensive measurements using a colored dye (rhodamine) and the factors regulating the change in dye
concentration, C(t) between measuring events are included in the Td model.

Td is first empirically determined from:

Td = -dt / ln (C(t) / C(t-1)) (2.17)

where
Td = the turnover (or retention) time (days);
dt = the time between measurements (days);
C(t) = the quantity of dye after the time (t) (lines);
C(t-1) = the quantity of dye at the previous measurement (liters).

The following regression gives Tdmod:

Tdmod = exp(7.72 - 2.93 · MFf - 29.26 / xm - 0.60 · ln(Ex)) (2.18)


(r2 = 0.82, n = 15, p < 0.001)

The mean filter factor (MFf in eq. 2.18) is an expression for how the coast outside the given coast
functions as an energy filter for the given coastal area; the filter factor was defined in fig. 1.63A, and
the mean filter factor is the filter factor (Ff in km3) divided by the number of openings. The smaller
the value for MFf, the denser the coast outside the coast, the more efficient the energy filter, and the
longer the deep water turnover time (Td).

The exposure (Ex) was discussed in sections 1.2.1 and 2.2. as a major determinant of the sensitivity of
coastal waters. The smaller the Ex value, the closer the opening towards the sea and/or surrounding
coastal areas, the smaller the wave energy impact in the given coastal area, and the longer the
theoretical deep water retention time, Td.

The mean slope (xm, in %) is defined according to Pilesjö et al. (1991). Very deep coastal areas with
large mean slopes are likely to have a larger volume of deep water and the theoretical deep water
retention time is likely to be longer than in coastal areas with smaller mean slopes.

Note that it is very difficult to transform these Tdmod values (fig. 2.31G) into a more normal frequency
distribution, because, there are many coastal areas with weak thermal stratification where it is difficult
to separate the deep water from the surface water and where the empirical Td values are set to 0. Some
Baltic coastal areas only mix once every spring and fall, and this means that the theoretical maximum
value for Td is about 120 days. Thus, there are no values for Td larger than 120 days in this dataset.

142
Also, note that the logarithmic transformation can NOT be used for Tdmod because log(0) is not
mathematically defined. Instead, log(l+Tdmod) is used.

The amount of material deposited in near-bottom sediment traps (1 m above the bottom: SedB in g
dw/(m2·day)). The traps were placed at 2-3 sites in each coastal area. They were out for about 7 days at
least 2 times during the period July to September in each coastal area. Obviously, the sedimentation of
material with an organic content of about 10-20 % influences the oxygen consumption, concentration,
and saturation. The higher the organic load, the lower the value for O2Sat (see also fig. 1.56 and fig.
1.61).

2. r-rank matrix. The aim of the following step is to provide a first screening of the factors
influencing O2Sat. Table 2.6 gives an r-rank matrix based on linear correlation coefficients, r (the r
value shows positive and negative relationships which the r2 values do not), for O2Sat versus different
variables assumed to influence O2Sat. High r values appear between O2Sat and some of the other
operational effect variables, especially Secchi depth (r = 0.68), sedimentation in near-surface
sediment traps (SedS, r = -0.65) and chlorophyll concentrations (r = -0.50). These high r values
indicate that good predictive models for O2Sat may be derived. It is interesting to note that TP and
TN both show high and significant negative r values versus O2Sat. -0.49 and -0.46; and that the
deep water retention time, Td, shows a strongly negative r value against O2Sat, r = -0.69.

Table 2.6. An r-rank table for O2Sat versus different effect, load and morphometric (sensitivity) variables.
A. Effect variables B. Load variables C. Sensitivity variables
O2Sat 1.00 Ntot -0.02 Fl 0.48
O2 conc 0.99 IP -0.16 BET 0.46
Secchi 0.68 Ptot -0.22 MFf 0.39
SedB -0.41 ANtot -0.22 Ff 0.36
Chl -0.50 APtot -0.30 A 0.27
SedS -0.65 IN -0.43 Ab 0.25
TN -0.46 At 0.10
TP -0.49 V 0.06
Dmax 0.04
Wb -0.02
E -0.08
Dr -0.26
Abbreviations are given in table 1.8 Ts -0.29
r values > 0.45 or < -0.45 (p < 0.05) Dm -0.33
for n=23 are bolded Xm -0.43
Ba -0.46
Vd -0.53
Td -0.69

3. Stepwise multiple regression. Stepwise multiple regression means that successively more model
variables (xi) are included in the regression, taking into account the interrelationships among the xi-
variables by means of their partial correlations coefficients. This means that variables from the same
clusters are not included in the model. Ideally, models of this kind should only contain variables
expressing different logical and mechanistically relevant functions and the xi-variables should emanate
from different clusters of related variables (or families). Stepwise multiple regression requires
advanced statistical software such as Statistica, SPSS, or the freely downloadable software R (www.r-
project.org).

143
Such a regression is presented in table 2.7. It should be interpreted like this: the "Step 1" row describes
the x-variable with the strongest explanatory power, x1, which in this case is log(TN+10·TP). The
equation including only y and x1 is given in the same row. The "Step 2" row describes x2, which is
the x-variable which adds the most explanatory power to the model in Step 1. In this case, it is log
(1+Tdemp). This procedure continues until there are no x-variables left which add any explanatory
power to the regression. Table 2.7 includes five x-variables, although it is often the case that fewer x-
variables than that are accepted at the 95% confidence level. The equation in the final step is the
model with the highest r2 value and will hereafter be referred to as "the O2Sat model".

Table 2.7. A "ladder" for a transformation of the operational effect variable O2Sat. The y-variable is
arcsin(O2Sat/100). F=4. Based on data from 23 coastal areas.
Step x-variable r2 Model
1 log(TN+10·TP) 0.49 y=15.6-5.38·x1
2 log(l+Tdemp) 0.76 y=13.14.39·x1-0.325·x2
3 log(Dm) 0.89 y=11.8-3.59·x1-0.351·x2-1.04·x3
4 Ff 0.91 y=13.2-4.15·x1-0.401·x2-0.905·x3-0.012·x4
5 log(V) 0.93 y=14.8-4.46·x1-0.403·x2-1.0263x3-0.021·x4+0.275·x5

One can note from Table 2.7 that O2Sat may be predicted quite well (r2 = 0.49) from the load
function (TN+10·TP). It is interesting that both nutrients are included in the model (and not just
nitrogen). Furthermore, the total concentrations of the nutrients (TN and TP) are used rather than
the inorganic fractions (IN and IP), although the latter are often promoted as more biologically
relevant in many eutrophication contexts. IN and IP are more reactive than TN and TP, which
implies that they vary more and that it is more difficult to access reliable collective data of these
variables. This can explain why IN and IP add much less to the predictive power of this model for
O2Sat than TN and TP. It is also interesting to note that the concentration variables (TN and TP) are
included in the model rather than the fluxes (Ntot, Ptot, ANtot and APtot). This is logical and
reflects the fact that the concentrations are the results of many internal processes regulating the
response of a given nutrient load to a given coastal area (internal loading, stratification, mixing, etc.).
The concentrations are good collective variables in this context. The most important morphometric
parameter is also logical, the mean depth (Dm), which reflects many aspects of the coastal character.
Ff and Td also entered the regression, indicating that the more sheltered the coast is, the longer the
theoretical deep water retention time, Td, and the lower the O2Sat. The r2 value is 0.93 for the
model (table 2.7) which indicates high predictive power (see section 2.4.3).

4. Clusters and functional groups. Predictive models should not only be built upon simple, easily
accessible, logical parameters. These parameters should also show a minimum of inter-dependence.
If two functionally related x-variables, such as Tdemp and Tdmod, are used in a model, their
variability will coincide so that variations in one model will convey very little more information
than the variation in the other has already added. A multiple regression or a dynamic model which
includes both Tdemp and Tdmod may discourage or eliminate the use of some other x-variable
which has the potential to add a considerable amount of new information to the regression. Analysis
of clusters (functional groups) is performed to eliminate the use of related variables. Since there is no
strict statistical definition of a functional group, a high r2 variable in a regression between two x-
variables is often used as an indication that they may belong to the same cluster or functional group.

5. Model range. Table 2.8 gives the range of the model variables. The O2Sat model should NOT be
used for coastal areas with characteristics outside these limits because the model has not been

144
tested for such conditions. Furthermore, the model cannot be used for coastal areas dominated by
tides, or for coastal areas from other coastal types than the ones from whence the data used to develop
the model comes from. If the model in table 2.7 is used for other coastal areas, then calculations must
be regarded with due reservations, as hypotheses rather than predictions.

Table 2.8. Ranges of model variables for the O2Sat model


TN TP TDemp Dm Ff V
(μg/l) (μg/l) (days) (m) (km3 (km3)
Min 256 14 0 3.8 0.059 0.0064
Max 417 31 126 13.8 30.7 0.18

6. Scatter plots. Fig. 2.34 gives six scatter plots (regression lines, r2 values, and p values for the 23
coastal areas) for empirical data on O2Sat versus some of the most interesting x-variables. From fig.
2.34A which gives the regression between O2Sat and the load function (TN+10·TP), one can note the
significant negative relationship (r2 = 0.38; p = 0.0019) and the rather even spread around the
regression line. It is evident that there must be a scatter since the empirical data for TN and TP
incorporate a certain uncertainty related to sampling and analysis and since many factors beside these
nutrieuts affect O2Sat. It is worthwhile, however, to note that as much as 38% of the variability among
the coasts in O2Sat can be statistically explained by this load function: the higher the nutrient load, the
higher the primary production and the higher the sedimentation of oxygen-consuming materials. The
next figure in fig. 2.34 gives the relationship between sedimentation in near-bottom sediment traps
(SedB) and O2Sat One can note a statistically significant (r2 = 0.17. p = 0.05) and logical negative
correlation. The variables presented in fig. 2.34 are targets in the following statistical analysis. All the
variables have also been transformed to increase the normality of the distribution as much as possible.

The negative relationship between O2Sat and the percentage of A-areas (BA; r2 = 0.21; p = 0.027) in
fig. 2.34C should also be noted: in coastal areas dominated a high wind/wave impact and a low
percentage of A-areas (and a high percentage of ET-areas), much of the oxygen-consuming material
will be transported rather quickly out of the coastal area and the values for O2Sat are high, and vice
versa Fig. 2.34D gives the negative relationship between O2Sat and the form factor; r2= 0.28, p =
0.009. In shallow coasts with a low form factor (fig. 2.34F), the oxygen consuming materials are
transported out from the coast, and deep coastal areas with a high form factor can act as efficient
sediment traps where the risks of getting low oxygen concentrations and low O2Sat values are high.
This is certainly also related to the theoretical deep water retention time, which is illustrated in fig.
2.34E.

7. Highest r2 and unexplained residual. One way of determining the highest possible degree of
statistical explanation of a predictive model is to compare two parallel empirical samples (re2), as
described in section 2.6. In this case-study, there are no such data available, but it is evident that a
significant part of the unexplained residual term [R = (1-r2) = (1-0.93) = 0.07] must be attributed to the
fact that the empirical data on O2Sat are somewhat uncertain. The coefficient of variation (CV) is
about 0.25 for O2Sat (see Wallin et al., 1992, and table 2.3). So, the highest reference r2 value
(rr2) is 0.96 (rr2 = 1-0.66·CV2), see eq. 2.8.

This means that the model presented in table 2.7 is very good indeed, under the given presuppositions.

145
Fig. 2.34. Scatter plots illustrating the relationship (regression line, r2 value and p value) between the operational
effect variable O2Sat (in %) and the load function (TN+10·TP), sedimentation in near-bottom sediment traps
(SedB), the percentage of A-areas (BA), the form factor (Vd), modeled values for the theoretical deep water
retention time (Tdmod) and the mean depth (Dm). n = 23.

8. Variants. The mean summer values of O2Sat include data from periods with different variability in
O2Sat. It is generally interesting in water management to predict mean summer O2Sat, since this is the
period when the lowest values are normally most likely to appear, but also other variants, like the
mean value from the winter period or values for defined months, may be of interest. In this section,
where the aim has been to link O2Sat to morphometric parameters, it was logical to use a mean value
from the longest possible registration period. In this case, no other variants of O2Sat have been
tested.

146
9. Outlier tests. The scatter plots given in fig. 2.34 are meant to reveal whether there are any outliers
(particularly conspicuous data points that should be omitted) in the data set. Since there are no clear
and interesting outliers identified in these scatter plots, further outlier tests have not been conducted.
There is a large scatter around all the regression lines in fig. 2.34, but this is the common case and has
to do with the fact that O2Sat is not highly related to any single load and/or sensitivity variable, but to
a combination of variables. All these model variables are based on extensive samplings, and there is
no reason to omit any data. All morphometric data are highly reliable (the error is generally lower than
CV ≈ 0.01; see Pilesjö et al., 1991). The uncertainties for the water chemical variables are higher, up
to about CV = 0.3 for IP and IN, CV = 0.1-0.2 for TN and TP (see table 2.3).

10. Confidence limits. Fig. 2.35 gives the relationship (regression line, the 95% confidence limits for
the predicted y and for the mean y) between model-predicted actual (untransformed) values and
empirical data of O2Sat. Note that this example describes an alternative model for O2Sat from
Håkanson (1999). It was developed from the same dataset but for a different purpose (see Håkanson,
1999), and the stepwise multiple regression yielded an r2 value of 0.87. The confidence limits for
individual y values are rather wide apart and the confidence bands for the mean y values quite narrow.
The r2 value of this model is 0.87 when regressed against empirical data, which is just as high a value
as was obtained in the multiple regression.

Fig. 2.35. Comparison between


empirical data on the operational
effect variable (actual value of O2Sat
in %) and model-predicted values
(model from Håkanson, 1999), and
95% confidence limits for the
predicted y and the mean y. Note that
the slope and the intercept of the
regression line are close to the ideal
line y = x. The r2 value is 0.87, the
same as for the multivariate regression
behind the model.

11. 3D-plots. Fig. 2.36 gives a 3D-diagrams where two of the model variables are given on the x- and
z-axes while the other model variables are kept constant The figure illustrate how this model predicts
O2Sat. The higher values of the nutrient load function, the longer the theoretical deep water retention
time, the lower the values of O2Sat (fig 2.36, which gives the model in table 2.7). The 3D-diagram
neatly illustrates how the model variables influence the target y-variable.

12. Conclusion. Many factors could potentially influence O2Sat in Baltic coastal areas. It is easy to
speculate and qualitatively discuss such relationships. With empirical data, it is possible to
quantitatively rank such factors and derive predictive models based on just a few, but the most
important, factors influencing O2Sat (for coasts of the given type). The tested working hypothesis -

147
"O2Sat depends on both load and sensitivity factors" - is supported by these results, which also
indicate which are the most important factors. Another tested hypothesis was: "It is possible to derive
practically useful empirical models". The given model could (statistically) explain about 90% of the
variability in the given y-variable among the 23 coastal areas. The highest reference r2 value (rr2) is
0.96. The load function and the theoretical deep water retention time are the two most powerful
predictors of O2Sat which is certainly mechanistically understandable.

Fig. 2.36. 3D-illustrations of


the model for O2Sat graph for
the model given in table 2.7.

A "natural" O2Sat could be estimated from the model if it is possible to estimate "natural" background
values of TN, TP and/or sedimentation (SedB). If the actual O2Sat value of the coast differs from such
a predicted "natural" value, then such divergences may be discussed in a quantitative manner. The
results are summarized in fig. 1.60.

Brief summary:
- Eutrophication and TP concentrations in lakes can be predicted with ELS models, using the TP
concentration in tributaries, and the lake water retention time.
- Water quality prediction in coastal areas is more complex than in lakes because the water exchange
with the sea exerts a great influence. This can partly be accounted for by means of morphometric
parameters.
- The oxygen saturation on deep bottoms is of fundamental importance to the benthic fauna which
needs oxygen to survive.
- O2Sat, a eutrophication indicator, could be predicted from load and sensitivity variables.
- Stepwise multiple regression is a powerful tool for making static ELS models.
- Regression techniques must be complemented by other procedures such as outlier analysis, cluster
analysis, transformation of variables, variant analysis, etc.
- A model is not valid outside its range, so it is important to document the range thoroughly.
- Regression plots with confidence limits and 3D plots are examples of how results can be displayed in
a way that makes good sense to other model users.

148
2.8. Model testing

This section mainly concerns dynamic ELS models, which are different than the static ELS models
described in the previous section in the sense that dynamic models describe changes over time,
using differential equations. Dynamic models are widely used in the aquatic sciences, and also in
many other scientific fields, from meteorology, soil science and geophysics, to medicine and
economics. Thus, the principles described here are not only relevant for this course and for its field of
application, but also for many other areas with which environmental analysts make contact.

The aim of this section is to give a few examples of two very used methods for critical model testing,
sensitivity tests and uncertainty tests. For further literature on this topic for aquatic ecosystem
models, see Håkanson and Peters (1995). The idea is not to promote a particular model - there are
other models available that are at least equally good as the ones tested here (see Bryhn, 2008 for a
review) - but instead to show how critical model testing is done in practice.

2.8.1. Calibration and validation


Before it is meaningful to test a model it should be calibrated and validated. Calibration means that a
given model set-up is tested against empirical data so that the fit between modeled values and
empirical data becomes as good as possible. One example of this is illustrated in fig. 2.37 using the
eutrophication model LEEDS for Lake S. Bullaren (fig. 2.3; see Håkanson, 1999 for information on
this lake and LEEDS) to test which value of the settling velocity (v) for particulate phosphorus
gives the best correspondence between modeled values and empirical data for TP concentrations
in water and surface sediments. If the v value is set too high, too much phosphorus will be deposited
in the sediments and too little will remain in the water, and vice versa. It is evident that a v value of
500 m/yr provides the best fit in this example.

All model variables (such as rates and distribution coefficients) could and should be tested like this,
and there are generally several combinations of values for the model variables that can give excellent
predictions when calibrated against a set of empirical data in one lake. All such combinations cannot
be correct if one seeks the model constants which could be used as default values in model
predictions for ecosystems in general. This means that a normal calibration involves iteration. The
idea with the calibration procedure is that for each round of iterations the uncertainty in the values for
the model variables should be reduced. When the model is duly calibrated, it should be validated, i. e.,
tested against independent data. It is evident that it is preferable if the calibrations and the validations
include as reliable empirical data from as many ecosystems as possible covering as wide a range
as possible in model variable characteristics.

To illustrate a validation and some inherent problems with models of the Vollenweider- and OECD-
type, fig. 2.38 gives a comparison between model-predicted CTP (using the OECD model) and
measured CTP for 18 lakes of varying limnological character in the domain of the OECD model (table
2.10). These lakes were not included in the development of the OECD model. The validation gives an
r2 of 0.45 for actual values and 0.4 for logarithmic values - i. e., rather poor results.

149
Fig. 2.37. Illustration of
calibration. Different
values for the fall velocity
(v) of particulate
phosphorus in Lake S.
Bullaren are tested in the
LEEDS model (see
Håkanson, 1999) to see
which v value can best
maintain the empirical
concentrations of
phosphorus in lake water
and sediments. The
empirical target value for
lake water is about 35 μg/l
and for surface lake
sediments, the target is
about 1.5 mg/g dw;
sediment concentrations
should lie between 1.8
(Max. emp. in fig. B) and
1.2 (Min. emp).

Fig. 2.38 also gives the 95% confidence intervals for the individual y for this regression. This
validation indicates that the uncertainty associated with the OECD model is so large that the model
is unsuitable for predicting CTP in individual lakes. Models of this type can give good predictions for
many lakes, even for most lakes (within the domain of the model), but - and this is important - it is
often not possible to predict for which lakes the models do and do not work! Basically, the models of
the Vollenweider type should not be used for lakes dominated by internal loading and changing
trophic conditions, but it may be difficult to develop operational criteria for this, and it is also
difficult to develop criteria for the level of statistical uncertainty that should be applied for the driving
variables, especially CTPin. How many samples are actually needed to determine the mean CTPin for
the tributaries to a given lake? It should also be noted that:

1. In spite of the fact that the model variables (Tw and CTPin) fall within the model domain, some of the
lakes used in this validation may have characteristics outside the range of the OECD model. For
example, the lake type may not be appropriate. The given validation may be incorrect for one or two
of the 18 tested lakes. This does not, however, affect the general conclusion that the OECD model is
too simplistic to provide meaningful predictions of lake TP concentrations, and hence also, lake
eutrophication effects, in many lakes.

2. Many of the lakes in table 2.9 probably have a significant internal loading of phosphorus (from
resuspension and diffusion). There is a close relationship between internal loading, bottom dynamic
conditions and the dynamic ratio (DR = √Area/Dm; see Håkanson and Jansson, 1983). If DR is larger
than about 2, more than 50% of the lake is likely to be dominated by areas of erosion and

150
transportation (ETareas). In such lakes, resuspension from wind/wave action is a most important
factor for lake TP concentrations.

Fig; 2.38. Illustration of a


validation of the OECD model
against an independent set of data
from 18 lakes.

Table 2.9. Lake data from an international data register. Tw = theoretical lake water retention time; Q = mean
annual water discharge; Dmax = maximum depth.; Dm = mean depth; DR = dynamic ratio; CTpin = mean annual
total-P concentration in tributaries; CTP = mean annual total-P concentration in lake water. OECD =mean annual
CTP calculated with the OECD model. Data from Meeuwig and Peters (1995).
Lake Area Vol Tw Q Dm Dmax DR CTPin CTP OECD
(106
(no) (km2) (m3·106) (yr) m3/yr) (m) (m) (-) (μg/l) (μg/l) (μg/l)
236 80.3 366 0.25 1464 4.6 6.4 1.97 77 38 39
241 1140 13600 2.2 6182 11.9 61 2.83 95 29 31
242 478 2900 3.3 879 6.1 22 3.60 74 41 23
243 1856 74000 55.9 1324 39.9 128 1.08 49 6 7
244 5648 153000 9 17000 27.1 106 2.77 52 10 13
246 867 8500 1.9 4474 9.8 60 3.00 25 10 11
247 1100 17800 2.7 6593 16.2 98 2.05 30 13 11
248 13.5 206 3.5 59 15.3 87 0.24 68 13 21
250 117.5 887 0.15 5913 7.5 36 1.44 36 34 23
261 1053 11900 0.25 4760 11.3 63.1 2.87 9 6 7
264 42.3 1640.1 13.4 122 38.8 84.1 0.17 49 16 11
271 3.8 55 0.47 117 14.5 30.2 0.13 26 1 14
272 7.4 118 0.47 251 15.9 29.3 0.17 12 4 8
277 11.5 1160 6.1 190 100.9 163 0.03 37 3 11
295 425 189.1 0.68 278 44.5 142.5 0.05 18 7 10
296 5.96 55.5 0.31 179 9.3 16.1 0.26 89 14 43
301 616 60000 10.6 5660 97.4 164 0.25 20 5 6
306 65.1 3300 1.1 3000 50.7 136 0.16 43 16 19
Ranges Min 0.1 1
for OECD Max 100 150
model variables

3. The high r2 value (0.86 for logarithmic TP concentrations) obtained in the derivation of the OECD
model can, probably, partly be attributed to the fact that the data used in that work emanate from a
scientific project where great efforts were made to ensure the quality of the data. Most data used in
practical water management are probably less reliable. This has strong bearings on the predictive

151
success. Many of the values in table 2.9, which emanate from an international data base, are, in all
likelihood, of rather poor quality. This is an often encountered problem, especially for many data-
bases available on Intemet and this problem should not be disregarded.

2.8.2. Sensitivity tests


Sensitivity analysis involves the study, by modeling and simulation, of how an alteration of one rate or
variable in a model influences a given prediction, while everything else is kept constant. This type of
analysis plays a dominant role in ecosystem modeling (see Hilton, 1993; Hamby. 1995; IAEA, 1998).
This section gives a typical example of how a sensitivity analysis can be performed. The calibration in
fig. 2.37 can also bee seen as a sensitivity analysis. The v value has been varied and all other factors
kept constant in a simulation to determine how variations in the given model variable influence a
given target variable. However, sensitivity analysis is a wider concept and it usually involves, at
least, two further steps.

Fig. 2.39 gives a sensitivity analysis where the v value from fig. 2.37 has been changed 100 times
while all other factors in the LEEDS model have been kept constant. In this case, it has been assumed
that there exist a typical, characteristic mean value for v (here 500 m/yr) and a given uncertainty for
this value given by a standard deviation, which has been set to 50% of the mean. That is, the CV value
bas been set to 0.5 for v. From a frequency distribution with a mean value of 500 and a standard
deviation of 250, 100 data have ken selected at random (by a generator in the software Ithink) and used
in the LEEDS model to produce the 100 curves for the target effect variable, PP (maximum volume of
phytoplankton). It is evident that the predictions of PP are very sensitive to the value selected for
the settling velocity (v).

Fig. 2.39. Sensitivity analysis using the LEEDS model in Lake S.


Bullaren. The target effect variable is PP and the settling velocity (v) has
been varied (mean value 500, CV = 0.5). 100 runs. The figure also
indicates that the following comparative sensitivity analyses use data for
the month yielding highest PP.

The next step in a sensitivity analysis is often to repeat this type of calculation for all interesting
model variables to try to produce a ranking of the factors influencing the target variable. The basic
idea is to identify the most sensitive part of the model, i. e., the part that is most decisive for the
model prediction. An example of such a comprehensive sensitivity analysis is given for the LEEDS
model in fig. 2.40.

152
Fig. 2.40. Results for sensitivity analyses using the LEEDS model in Lake S. Bullaren.

In this case, the following model variables were included in the test the tributary TP concentration
(Cin, in fig. A), total TP-emissions from a fish cage farm producing 500 ton/yr of rainbow trout (B),
the distribution coefficient for lake water regulating the TP fluxes into dissolved and particulate
fractions (Kd, fig. C), the settling velocity (v, fig. D), the concentration of suspended particulate
matter in water (SPM, fig. E) and, finally, the theoretical lake water retention time (Tw, fig. F). These
sensitivity analyses have been done for the lake TP concentration (lower curves in fig. 2.40) and the
effect variable PP (upper curves in fig. 2.40). In these sensitivity analyses all model variables were
altered by the same factor (1.5). From fig. 2.40, one can note that from these presuppositions, the TP
and PP predictions are most sensitive to the choice of the values for Cin, Kd and v. This is certainly
very logical and applies for many substances in many types of aquatic ecosystems because these
three model variables regulate two primary fluxes, inflow (Cin) and sedimentation (Kd and v). All
other fluxes depend on these fluxes. Fig. 2.41 gives the actual time-series of data for fig. 2.40A.
Such time-series must be calculated for all the given model variables before the results can by
summarized in the manner given in fig. 2.40.

However, it is evident that it is NOT realistic to apply the same uncertainty for all model
variables, like a factor of 1.5 in the previous example. There are major differences among model
variables in this respect. Morphometric parameters can often be determined very accurately (see
Pilesjö et al., 1991); some model variables, like rates and distributions coefficients, can, on the other
hand, not be empirically determined at all for real ecosystems, but have to be estimated from
laboratory tests or theoretical derivations. This means that the values used for such model
variables are often very uncertain. Table 2.10 gives a compilation of typical, characteristical CV
values for many types of variables used in ecosystem models. From this table, one can note that model
variables like rates and distribution coefficients generally can be given CV values of 0.5. The highest

153
expected CV values appear for certain sedimentological variables, like TP concentration in T- and E-
sediments. In the following uncertainty tests, we will use the CV values given in table 2.10.

Fig. 2.41. Sensitivity analyses for


the tributary TP concentration
(Cin) in the LEEDS model for
Lake S. Bullaren for the target
effect variable, PP (A), and lake
TP concentrations (B).

2.8.3. Uncertainty tests using Monte Carlo techniques


Two main approaches to uncertainty analysis exist, analytical methods (Cox and Baybutt, 1981; Beck
and Van Straten, 1983; Worley, 1987) and statistical methods, like Monte Carlo techniques (Tiwari
and Hobbie, 1976; Rose et al., 1989; IAEA, 1989). In this section, we will only discuss Monte Carlo
simulations, which are based on generated random values series with user-specified CV values.

Uncertainty tests using Monte Carlo techniques may be done in several ways, using uniform CV
values, or more realistically, using characteristic CV values (e. g., from table 2.10). For predictive
empirical or dynamic ELS models based on several uncertain model variables (rates, etc.), the
uncertainty in the prediction of the target variable (y or E) depends on such uncertainties. The
cumulative uncertainty from many uncertain x-variables may be calculated using Monte Carlo
simulations, and that is the focus of this section.

Monte Carlo simulations is a technique to forecast the entire range of likely observations in a given
situation; it Can also give confidence limits to describe the likelihood of a given event. Uncertainty
analysis (which is a term for this procedure) is the same as conducting sensitivity analysis for all
given model variables at the same time. A typical uncertainty analysis is carried out in two steps.
First, all the model variables are included with defined uncertainties and the resulting uncertainty
for the target variable calculated. Then, the model variables are omitted from the analysis one at
the time. The procedure is illustrated in fig. 2.42.

154
Table 2.10. Compilation of characteristic CV values for different types of lake variables. All CV values, except
for the model variables, emanate from empirical measurements.
CV CV
Catchment variables Lake management variables
Catchment area (ADA) 0.01 Secchi depth (Sec) 0.15
Percent outflow areas (OA) 0.10 Chlorophyll-a concentration (Chl) 0.25
Fallout of radiocesium (Cssoil) 0.10 Hg concentration in fish muscle 0.25
Mean soil type or permeability factor (SP) 0.25 Cs concentration in fish 0.22
Cs concentration in water 0.30
Lake variables Cs concentration in sediments 0.60
Lake area (Area) 0.01
Mean depth (Dm) 0.01 Climatological variables
Volume (Vol) 0.01 Annual runoff rates 0.10
Maximum depth Dmax) 0.01 Annual precipitation 0.10
Theoretical water retention time (Tw) 0.10 Temperatures 0.10

Water chemistry variables Model variables


pH 0.05 Fall velocities 0.50
conductivity (cond) 0.10 Age of A sediments 0.50
Ca concentration (Ca) 0.12 Age of ET sediments 0.50
Hardness (CaMg) 0.12 Diffusion rates 0.50
K concentration (K) 0.20 Retention rates 0.50
Colour (col) 0.20 Bioconcentration factors 0.50
Fe concentration (Fe) 0.25 Feed habit coefficients 0.50
Total-P concentration (TP) 0.35 Distribution (=partition) coefficients 0.50
alkalinity (alk) 0.35

Sedimentological variables
Percent ET-areas (ET) 0.05
Suspended particularer matter conc (SPM) 0.20
Mean water content for E-areas 0.30
Mean water content for T-areas 0.20
Mean water content for A-areas 0.05
Mean bulk density for E-areas 0.10
Mean bulk density for T-areas 0.10
Mean bulk density for A-areas 0.02
Mean organic content for E-areas 0.50
Mean organic content for T-areas 0.50
Mean organic content for A-areas 0.10
Mean TP-conc. for E-areas 0.50
Mean TP-conc. for T-areas 0.75
Mean TP-conc. for A-areas 0.35
Mean metal conc. fot E-areas 0.50
Mean metal conc. for T-areas 0.75
Mean metal conc. for A-areas 0.35

Figures 2.43 and 2.44 give such an uncertainty analysis for a version of the classical ELS model (see
fig. 2.45) using the basic mass-balance model for TP and the regression between CTP and PP for data
from Lake S. Bullaren, Sweden. The CV values from table 2.10 have been used in this test. These two
figures also give results for sensitivity analyses, which can be directly compared to the results for the
uncertainty analyses. One hundred runs have been done and the variabilities (including CV) for both
CTP and PP have been determined for the year yielding maximum PP values.

One can note that the most crucial uncertainty component for CTP and PP is the value used for Cin, the
TP concentration in the inflow. The values selected for the settling velocity (v), the exponent for the
retention rate (1/Twexponent) or the specific runoff rate (SR) influence TP and PP predictions much

155
less. The uncertainties associated with the morphometric data (CV = 0.01) do not affect the predictions
in any significant manner. The figures also give the calculated CVs for the target variables, CTP and
PP. These CVs can be used to rank the influence that the model variables have, under the given
conditions, on predictions of CTP and PP.

Fig. 2.42. Illustration of the principles of


uncertainty analysis using Monte Carlo
simulations.

Another aspect of this uncertainty test concerns the value for the settling velocity v for TP. In the
classical ELS model, v is set to 5 m/yr as a default value (see fig. 2.45). This is a calibrated value for
which Lake S. Bullaren gives the best predictions of the empirical lake TP concentration of about
35 μg/l. This value can be compared to the value used for v in the comprehensive LEEDS model,
which is 500 m/yr, a difference of 100. Both values cannot be correct as generic values for the
settling velocity for phosphorus in lakes! The value v = 5 is a typical lake-specific variable and it
has to be changed for every lake after careful calibrations against lake-specific empirical data. If this
value is used generally, the model will yield poor predictions. The value of 500 m/yr in LEEDS, on
the other hand, is valid for particulate-P, the only fraction of phosphorus that can settle out in lakes,
and it is included in a model structured in such a manner that it could be a used for all lakes. That is, v
in LEEDS is supposed to be a model constant to be used in order to avoid the problems with lake-
specific model tuning that were discussed in 2.1.

The default settling velocity for particulate radiocesium is lower than for particulate phosphorus; only
15 m/yr. Radiocesium is mainly associated with clay minerals of the illite-type, which settle relatively
slowly in lakes. Particulate phosphorus, on the other hand, is associated with seston (i. e., dead

156
plankton), large flocs of Fe-Mn-oxides and hydroxides and various types of suspended and
resuspended inorganic and organic materials.

Fig. 2.43. A. Uncertainty analyses using the classical ELS model (see fig. 2.45) for TP-concentrations in Lake S.
Bullaren. 100 runs. The box-and-whisker plots show the median (M50), quartiles, percentiles and outliers. CV
values for the model variables are given as well as calculated CV values for the TP-concentration.
B. Corresponding results from sensitivity analyses.

Fig. 2.46 illustrates the relationship between the settling velocity and the grain size according to
Stokes' law. Carrier particles of clay-size have relative small v values, in the order of about 10 m/yr.
Most particles of silt-size have fall velocities in the range from about 100 to 1000 m/yr.

The results given in figures 2.43 and 2.44 are indicative of a typically poorly balanced model since
the variability in the target variables (TP and PP) is so highly dependent on the variability in one of
the driving variables (Cin). The calculated uncertainties for these 100 runs for the target variables TP
and PP are given in the figure, and together with the box-and-whisker plots, they demonstrate that this

157
model and all models of the Vollenweider- and OECD-type, and hence, all classical ELS model for
lake eutrophication, are poorly balanced and highly dependent on the reliability of the data for Cin.

Fig. 2.44. A. Uncertainty analyses using the classical ELS model for PP in Lake S. Bullaren. 100 runs. CV
values for the model variables are given as well as calculated CV values for PP. B. Corresponding results from
sensitivity analyses.

This is nothing new. It is actually the reason why so much effort (see, Dillon and Rigler, 1974, 1975;
Nichols and Dillon, 1978; Chap and Reckhow, 1979,1983) have been devoted to developing a good
understanding of the processes regulating the tributary flux of phosphorus and the efforts to identify
sources for TP-fluxes from catchments. This is the key factor behind poor predictions with respect
to lake eutrophication. A well balanced model should NOT be too dependent on the uncertainty
related to a single variable. Instead, all the box-and-whisker plots should look alike and the CV
values for y should NOT change too much if a model variable is omitted in the uncertainty analysis.

An interesting and paradoxical result in Monte Carlo simulations of dynamic models can be obtained
if too few runs are made. Then the uncertainty in the target variable can increase when model
variables are eliminated in the uncertainty test. This contradicts what can be analytically shown

158
using, e. g., the Gauss approximation formula (see, e. g., Blom. 1989; Jonsson, 1998). It is evident that
the more runs, the better the results of the Monte Carlo simulations. If few runs are being made,
occasional random outliers can disrupt the overall picture and create erroneous CV values for the
calculated target variables. From figures 2.43 and 2.44, one can note that occasional outliers cause
totally unrealistic TP concentrations and PP values in Lake S. Bullaren, where the empirical TP
values is about 35 μg/l and the corresponding PP value about 3 mm3/l.

Fig. 2.45. The "classical" ELS model.


Equations, target variables, lake-specific
variables and model variables.

The importance of the certainty of Cin, as demonstrated in fig. 2.43 and 2.44, is further exemplified
for PP-predictions using the LEEDS model in fig. 2.47. The figure shows Monte Carlo simulations
when CV for Cin has been set to 0.5. This uncertainty analysis involves 15 model variables in LEEDS.
Most of them are defined in fig. 2.48 and the CVs used are also given in the figure. Fig. 2.47 gives the
results when the uncertainties of all 15 model variables are accounted for. This gives a calculated CV
for PP of 54%, which is high. Occasional outliers give PP values higher than the alarm limit of 10
mm3/l.

The four box-and-whisker plots in fig. 2.47 were obtained when Cin, v, epilimnetic water temperatures
(temp) and the distribution coefficient (Kd) were omitted from the uncertainty analysis. These are the
four most important model variables according to this test. The main message from fig. 2.47 is that
LEEDS is a well-balanced model. Cin is, however, still the most important model variable for the

159
uncertainty in the PP-predictions, but CV for PP does not decrease more than 14.6 percentage units
(from 54.0% to 39.4%) when Cin is omitted.

Fig. 2.46. The relationship


between the settling velocity (v
of spherical particles) in water,
particle diameter and particle
density (typical values for
humus is = 1.5, clay = 2.4,
quarts - 2.7, dolomite = 2.8) as
given by Stokes's law (at 20°C).

This means that it is very important to have a reliable characteristic CV value for Cin. Fig. 2.49
gives a frequency distribution for 98 measurements of TP concentrations in tributaries to Lake S.
Bullaren (from Håkanson and Johansson, 1995). The median CV value is 0.5, which is used as default
CV for Cin in these simulations. This value may be typical just for these tributaries this year (1994). It
is, however, likely that CV for Cin is generally greater than CV for C, since the seasonal variations
in TP concentrations are likely to vary more in rivers than in lakes. A characteristic CV for C is 0.35,
so it is probable that a characteristic CV for Cin for Nordic rivers is about 0.5.

Fig. 2.47. Monte Carlo


simulations using the LEEDS
model for Lake S. Bullaren.
The box-and-whisker plots
show the results for the four
model variables contributing
the most to the uncertainty in
the target variable, PP. The
characteristic CV values are
given as well as the calculated
CVs for PP.

160
Fig. 2.48. Monte Carlo simulations using the LEEDS model in Lake S. Bullaren without uncertainties associated
with Cin, the main contributor to the overall uncertainty in the target variable, maximum volume of
phytoplankton (PP).

To conclude: The uncertainty in Cin is the most important factor in the LEEDS model as well, and
when this uncertainty is omitted the uncertainty in PP drops from 0.54 to 0.39.

Fig. 2.3 in section 2.1.3 gives a compilation of all major P fluxes in Lake S. Bullaren according to
simulations with the LEEDS model. One can note that the typical tributary inflow is about 4000 kg
P/yr and the total inflow from the fish farm about 1900 kg P/yr. These are the two most important TP
loading fluxes. All other fluxes depend on these two.

Fig. 2.49. The frequency


distribution for the 98 empirical
measurements of CTPin (Cin in
the figure) in tributaries to Lake
S. Bullaren, 1994. Modifled
from Håkanson and Johansson
(1995).

2.8.4. Uncertainty and sensitivity analysis as tools for structuring ELS models
An interesting and powerful application of Monte Carlo simulations concerns the use of this method to
identify important modeling structures and, hence, Monte Carlo methods can be used as a tool for
building predictive ELS models. If one first identifies the most important factor regulating the
predictions of a target variable, like Cin for PP, the next step is to omit Cin from the uncertainty
analysis and re-iterate the Monte Carlo simulations. This means that the uncertainty associated with
this variable is removed and the default value is used in the following Monte Carlo simulations (like

161
Cin = 60 μg/l for Lake S. Bullaren). Then, the Monte Carlo simulations can be used to identify the
second most important modeling variable (and the third, fourth, fifth, etc.).

Fig. 2.50. Monte Carlo simulations


using the LEEDS model for Lake S.
Bullaren for different CV values for
Cin (0.5, 0.35, 0.2) and for a
situation when the uncertainties
associated with Cin are neglected.

The method presupposes that characteristic CV values are used for all model variables, and NOT
uniform CVs. This procedure is illustrated in fig. 2.48. The first box-plot describes Monte Carlo runs
with all uncertainties (except that of Cin) included. In the other box-plots, the uncertainty of one
variable at a time has been eliminated (see fig. 2.42). Thus, fig. 2.48 shows that the second most
important model variable for PP in Lake S. Bullaren using the LEEDS model is the settling velocity
(v). Omitting uncertainties associated with v reduces the calculated CV for PP from 39.4% to 29.4%.
Moving down the ranking list, the third most important model variable for PP is the dimensionless
moderator Ybio, which regulates how the light conditions, surface water temperature and water clarity
influence PP. The fourth most important model variable is Kd, the distribution coefficient which
regulates how much phosphorus is in dissolved and particulate phases. All other model variables
contribute less to the overall CV for PP.

Figure 2.48 also gives characteristic CVs for the given model variables. One can note that the
uncertainties associated with the mixing rate (Mix), the mineralization rate (Min) and the diffusion rate
(Diff) contribute relatively little to the CV for PP. This is also the case for all the model variables
associated with the fish farm; the feed conversion ratio (FCR), the TP concentration in feed (Pfeed)
and in the cultivated rainbow trout (Pfish).

A complementary comparative sensitivity analysis for the second, third and fourth most
important model variables (v, Ybio and Kd; identified in fig. 2.48) and the two least important
model variables (Mix and Diff) is given in fig. 2.51. One can note that the sensitivity with respect to
the goal variable (PP) is rather similar for the five variables, although v yields the greatest sensitivity,
followed by Ybio.

Thus, box-and-whisker plots in the uncertainty and comparative sensitivity analyses of LEEDS
(figures 2.47, 2.48, and 2.51), and all the associated CVs for PP in each of fig. 2.47 and 2.48, are also

162
quite similar. This, again, confirms that the LEEDS model is a well-balanced model. Results in these
three figures are rather different from the results given in fig. 2.43 regarding the classical ELS model.

Fig. 2.51. Sensitivity analysis using the LEEDS model in Lake S. Bullaren for the three most important model
variables and the two least important model variables identified in the uncertainty analysis in fig. 2.48. When the
uncertainties associated with Cin are omitted, the main contributor to the overall uncertainty in the target
variable, maximum volume of phytoplankton, PP, is the settling velocity.

Brief summary:
- The main idea with uncertainty and sensitivity tests is to identify the most important weaknesses of
the model.
- When the weakest parts of the model have been identified, one should focus future work on
strengthening these parts.
- The idea is also to omit unnecessary components of the model that add uncertainty but no predictive
power. The simpler, the better!
- Sensitivity analysis is a method for analyzing the uncertainty in the goal variable from one model
variable (x) at a time, while ignoring the uncertainty in the other model variables.
- Uncertainty analysis is to study the uncertainty in the target variable from uncertainties in all model
variables simultaneously, and comparing the target variable uncertainty from this case to cases during
which the uncertainty in one model variable at a time is omitted.
- The certainty of the average phosphorus concentration in the tributaries is the single most important
determinant of the predictive success of phosphorus models for lakes.

163
3. Epilogue
Differences in aquatic ecosystems can be manifested in many ways, by many different aspects of
biology, water chemistry and morphometry. All lakes and coastal areas are therefore individual
ecosystems, but behind those many individual manifestations, there are structural and functional
similarities. The basic questions that this book asks are:

- How can different chemical threats to aquatic systems be quantitatively ranked?


- Which processes regulate spread, biouptake and effects in aquatic ecosystems of chemical
pollutants?
- How can the most important environmental problems be modeled and remediated?

This textbook is multi-disciplinary in the sense that biological, chemical, geological, geographical,
statistical and mathematical terms and concepts have been used, and that different types of variables,
models and processes have been discussed. However, this book is also specific in the sense that the
focus is on lakes and coastal areas and at the ecosystem scale.

The PER approach is meant to be used as a tool to highlight what we know about chemical threats to
real aquatic ecosystems and to give an avenue to pose questions in a structured manner towards a
defined goal. ELS modeling may be regarded as the "hub of the PER wheel". ELS models should be
simple. They can be used to simulate the likely ecosystem effects of remedial measures.

To be practically useful, ELS models must yield high predictive power for the defined target effect
variable. Ideally, such target effect variables express a change in toxin concentration, abundance,
reproduction or biomass of defined key functional organisms, preferably at the highest trophic level.
Very often it is not possible to reach that goal, but in the PER analysis it is always important to keep
that goal in mind and to be open about how far from the goal the "state-of-the-art" for a given
contaminant is. Generally, for practical and economic reasons, one must seek operational effect
variables which can be easily measured and modeled in contexts of water management. Some
examples of such target variables exemplified in this book are O2Sat, pH, the Secchi depth, Hgpi and
Cspi.

One can easily find limitations and flaws in the given PER approach and the presented ELS models.
We would therefore like to challenge our readers to use their criticism to improve the methods
presented in this book, or to develop alternative methods that can do the same trick but in an even
more robust manner. As Karl Popper wrote, seeking the truth in order to get as close to it as possible is
"one of the strongest motives for scientific discovery". This quest is pursued by means of repeatedly
testing and improving theories, methods and models. Furthermore, this quest is not only essential to
scientists, but to all practically inclined environmental analysts and managers that are interested in
understanding and combating prevailing and future problems in the aquatic environment.

Table 2.4 presented a compilation of ELS model achievements regarding r2 values in mod-emp
regressions compared to highest possible r2 values for ELS models. There is much improvement left to
be done according to this table, and we anticipate that corresponding tables in the future will convey
more and higher numbers. Our pious hope is that some of the readers of this book will take part in this
endeavor, and that they and others will find the facts and methods in this book practically useful.

164
Literature references

Most of this book is based on the first two chapters in

Håkanson, L., 1999. Water pollution - methods and criteria to rank, model and remediate chemical
threats to aquatic ecosystems. Backhuys, Leyden, 277 p.

All references from 1999 and before are listed in Håkanson (1999) and can be downloaded at
http://www.geo.uu.se/miljoanalys/pdf/waterpollutionreferences.pdf

Additional references are:

Bryhn, A. C., 2008. Quantitative understanding and prediction of lake eutrophication. PhD thesis,
Uppsala University, 38 p + 5 thesis articles.

Håkanson, L. and Bryhn, A. C., 2008. Tools and Criteria for Sustainable Coastal Ecosystem
Management. Springer, 2008, 292 p.

165

View publication stats

Вам также может понравиться