• No results found

Appropriate Modelling Complexity: An Application to Mass Balance Modelling of Lake Vänern, Sweden

N/A
N/A
Protected

Academic year: 2022

Share "Appropriate Modelling Complexity: An Application to Mass Balance Modelling of Lake Vänern, Sweden"

Copied!
48
0
0

Loading.... (view fulltext now)

Full text

(1)

Comprehensive Summaries of Uppsala Dissertations from the Faculty of Science and Technology 971

Appropriate Modelling

Complexity: An application to mass-balance modelling of Lake

Vänern, Sweden

BY

M AGNUS D AHL

ACTA UNIVERSITATIS UPSALIENSIS

UPPSALA 2004

(2)

Dissertation at Uppsala University to be publicly examined in 9C204 (Ericssonsalen), Karlstad University, Wednesday, May 26, 2004 at 11:00 for the Degree of Doctor of Philosophy. The examination will be conducted in Swedish.

Abstract

Dahl, M. 2004. Appropriate Modelling Complexity: An application to mass-balance modelling of Lake Vänern, Sweden. Acta Universitatis Upsaliensis. Comprehensive Summaries of Uppsala Dissertations from the Faculty of Science and Technology 971. 42 pp. Uppsala. ISBN

91-554-5950-1

This work is about finding an appropriate modelling complexity for a mass-balance model for phosphorus in Lake Vänern, Sweden. A statistical analysis of 30 years of water quality data shows that epilimnion and hypolimnion have different water quality and should be treated separately in a model. Further vertical division is not motivated. Horizontally, the lake should be divided into the two main basins Värmlandssjön and Dalbosjön. Shallow near shore ares, bays and areas close to point sources have to be considered as specific sub-basins if they are to be modelled correctly.

These results leads to the use of a model based on ordinary differential equations. The model applied is named L

EEDS

(Lake Eutrophication Effect Dose Sensitivity) and considers phosphorus and suspended particles. Several modifications were made for the application of the model to Lake Vänern. The two major ones are a revision of the equations governing the outflow of phosphorus and suspended particle through the outflow river, and the inclusion of chemical oxygen demand (COD) into the model, in order to model emissions from pulp and paper mills. The model has also been modified to handle several sub-basins.

The L

EEDS

model has been compared to three other eutrophication models applied to Lake Vänern. Two were simple models developed as parts of catchment area models and the third was a lake model with higher resolution than the L

EEDS

model. The models showed a good fit to calibration and validation data, and were compared in two nutrient emission scenarios and a scenario with increased temperature, corresponding to the green house effect.

Keywords: Lake Vänern, phosphorus, mass balance modelling, model complexity

Magnus Dahl, Department of Earth Sciences. Uppsala University. Villavägen 16, SE-752 36 Uppsala, Sweden

 Magnus Dahl 2004 c

ISBN 91-554-5950-1 ISSN 1104-232X

urn:nbn:se:uu:diva-4239 (http://urn.kb.se/resolve?urn=urn:nbn:se:uu:diva-4239)

(3)

List of papers

The thesis is based on the following papers, which in the text will be referred to by their Roman numerals:

I. Magnus Dahl and David I. Wilson, July 2002, The danger of short-term validation for lake models. Submitted.

II. Magnus Dahl and David I. Wilson, 2004, Is Lake Vänern well mixed? A statistical procedure for selecting model structure and resolution. Jour- nal of Great Lakes Research 30(2). In press.

III. Magnus Dahl and David I. Wilson, 2003, Modelling salt transport in Baltic basins. In proceedings of the IASTED International Conference on Modelling, Identification, and Control, February 10–13, 2003, Inns- bruck, Austria.

IV. Magnus Dahl, David I. Wilson, and Lars Håkanson, 2003, A combined suspended particle and phosphorus water quality model: Application to Lake Vänern. Submitted.

V. Magnus Dahl and B. Charlotta Pers, Comparison of four models simu- lating phosphorus dynamics in Lake Vänern, Sweden. Submitted.

My part of the work in papers I, III, and IV concerns all the simulations

and model modifications, along with part of the writing. In paper II, I have put

together and analysed the data and done the major part of the writing and in

paper V I have done the L EEDS and F YRISÅ model simulations and about half

the writing.

(4)

Contents

1 Introduction and aim . . . . 1

2 Study area – Lake Vänern . . . . 1

3 Lake models . . . . 4

4 What’s an appropriate model for Lake Vänern? . . . . 5

4.1 The spatial resolution justified by data (II) . . . . 5

4.2 How to model mixing between basins (III) . . . . 9

5 Application of the L EEDS model to Lake Vänern . . . 12

5.1 The need of long data series for calibration (I) . . . 13

5.2 Model modifications for Lake Vänern (IV) . . . 14

5.3 Running the L EEDS model with 1, 2, and 5 sub-basins (IV) . . . . 18

6 Comparison of eutrophication models for Lake Vänern (V) . . . 19

7 Conclusions . . . 25

Description of the L EEDS model . . . 27

Acknowledgements . . . 35

Summary in Swedish . . . 36

(5)

1 Introduction and aim

Many lakes are, or have been, polluted by industrial and municipal wastes.

Computer simulation is a useful tool to determine the amount of remedial ac- tions needed to restore the lake to the desired water quality. There are further benefits of the better ecosystem knowledge provided by a model. Many dif- ficult management questions, regarding for example maximum allowed emis- sions from an industry or the possible need to reduce nutrient runoff from farmland, would be easier to resolve with the information that a water quality model can give. Unfortunately, water quality modelling is less mature than many other applications of computer simulation. No standard ecosystem mod- els exist the way they do in for example mechanics or chemical engineering, with well known precision and application areas. The reason is the high cost and low accuracy and precision of ecological data along with the large com- plexity of ecosystems. There is also a need to keep a sampling program run- ning for several years with at least monthly samples to collect enough data for model calibration. Lake Vänern, Sweden, has an unusually long data series spanning the last 30 years; for a few variables even the last 100 years, and is therefore well suited for examination of models.

The aim of this thesis is to examine what is an appropriate model for long term effects of emission of nutrients, mainly phosphorus, to Lake Vänern, to develop such a model, and to apply it to some representative scenarios.

Nitrogen is not included in the model, which might be seen as a disadvantage,

as the nitrogen in Lake Vänern contributes to the eutrophication once it reaches

the sea through River Göta Älv, and predictions of the nitrogen reductions from

different remedial actions are therefore desired. From the perspective of an

ecosystem model for Lake Vänern, however, nitrogen is not important, as the

bioproduction is limited by phosphorus. Unimportant variables should not be

included in a model, because they contribute to the uncertainty of the model

without adding any extra information (Håkanson and Peters, 1995). Therefore,

nitrogen is not included in this study, but it has been in other model studies

of Lake Vänern (Sonesten, In prep.; Pers and Persson, 2003; Arheimer and

Brandt, 1998).

(6)

2 Study area – Lake Vänern

Lake Vänern (figure 1) is the largest lake in Sweden, containing one third of Sweden’s fresh water. With an area of 5,648 km 2 and a volume of 153 km 3 it is also the third largest lake in Europe, after the Russian lakes Ladoga and Onega. The maximum depth is 106 m, the average depth 27 m, and the resi- dence time is 9 years (Statens Naturvårdsverk, 1978). The catchment is 46879 km 2 and consists mainly of forests (52%), lakes (19%), and farmland (12%) (Wallin, 1994). The largest rivers enter from the north, and have humic and nutrient-poor water from forests and mountains. The smaller rivers on the south and south-west come from farmland and have higher nutrient concentra- tions (Statens Naturvårdsverk, 1978).

Lake Vänern is oligotrophic, and especially the phosphorus concentration is low (Wilander and Persson, 2001). As a consequence there is also little algae, plankton, fish, and plants in the lake. The lake is also dimictic, which means that the bottom and surface water mix twice a year, at spring and autumn, and that there is generally ice cover in the winter and a temperature stratification in the summer, with light warm water on top of cold heavier water, separated by a rather sharp temperature gradient (Statens Naturvårdsverk, 1978).

The area around Lake Vänern is one of the centra for pulp and paper industry in Sweden. The emissions of organic matter and nutrients to the lake started to have noticeable effects on the water quality about 1910, and reached a peak in 1965, when 75% of the supplied organic matter came from the industry. The industrial load has now decreased to 17% of the total organic load (Wallin, 1996), making the present organic matter concentration in the lake the lowest since 1910. The use of elementary chlorine free bleaching of pulp has resulted in drastically decreased emissions of chlorinated organic matter, and the emis- sions of mercury from a chlor-alkali factory have stopped (Lindeström, 1995, 2001). The main sources of nutrients today are cities, farmland, and pulp and paper mills. The cities and pulp mills built waste water treatment plants during the 1970s (Wilander and Persson, 2001), but the emissions from farmland and atmospheric deposition have proved more difficult to reduce. Pollution from other sources is low, except the large emissions of zinc from a rayon factory (about 70 % of the inflow of zinc in the 1960s). It has now decreased to 10%

of its maximum value (Wallin, 1996).

In the 1960s, when the lake was at its most polluted state, a project was initiated

to investigate the lake, the water quality, and to produce a mathematical model

to simulate cleanup scenarios. The project was carried out between 1972 and

1977 and resulted in two circulation models and a mercury model, but not the

(7)

59°20’

59°00’

58°40’

58°20’

14°00’

13°00’

Karlstad

Kristinehamn

Mariestad

Lidköping

Vänersborg Åmål

Säffle Grums

Skoghall

2 3

1

6 4

9

8

10

7 5

0 10 20 30 40 50 km

Göta Älv (outflow)

Dalbosjön

Värmlandssjön

Ν

Pulp mill Still used Used until 1995 Used until 1978 Used rarely Sampling locations

Figure 1: A map of Lake Vänern, including the sampling locations and the pulp and paper mills adjacent to the lake.

ecosystem model intended. The reasons for this were mainly the absence of knowledge in the field of ecosystem modelling, and a lack of relevant data for Lake Vänern. Valuable results from the project include a report summarising the investigations (Statens Naturvårdsverk, 1978), and a monitoring program (samples taken once per month at ten representative locations in the lake). The monitoring program is still in operation, but the number of sampling locations was reduced to three in 1995.

In the middle of the 1990s, phosphorus and COD had approached their natural

levels (Sonesten, In prep.; Wallin, 1996). Despite this, further reduction of

phosphorus and COD was prescribed by authorities according to a BAT – best

available technology – policy. The waste water treatment plants are expensive

to build and run, and harm the environment due to energy consumption and

production of sludge. The energy consumption contribute to the green house

effect and acid rain. The sludge contains pollutants precipitated from the water,

and is normally transported to an incinerator and burned at high temperature,

with negative effects on the environment. Again, an ecosystem model of Lake

(8)

Vänern became the object of discussion, and this modelling project started.

Today, there is 30 years of data from the measurement program initiated by the 1970s project (easily accessible from http://info1.ma.slu.se/). There is also much better software for modelling, and many models exist to study.

All the data used in this project are collected as part of standard measurement programs. Monthly values of water chemistry and temperature from within the lake and from the main rivers are achieved from a data base maintained by SLU, the Swedish University of Agricultural Sciences (http://info1.ma.

slu.se/). Data on river flow in the main rivers, water level, and weather (air temperature, rainfall etc.) comes from SMHI, The Swedish Meteorological and Hydrological Institute. Data on point source emissions to the lake are taken from the TRK project (Brandt and Ejhed, 2002), (http://www-nrciws.

slu.se/TRK/), from Wallin (1994), and in some cases from the industries and municipalities in question.

More information about the lake can be found at Vänerns Vattenvårdsförbund (www.vanern.s.se), at the Lake Vänern museum in Lidköping, and from Am- bio Vol. 30, No. 8, Dec. 2001, which is a thematic issue on the four large lakes of Sweden.

3 Lake models

A review of lake phosphorus models, the modelling procedure, and useful soft- ware has been done as a part of this project (Dahl and Wilson, 2000). A short summary of the report is given here. Part of the results have also been pub- lished in a conference proceedings (Dahl et al., 2001).

The simplest phosphorus models for lakes are retention models, stating how much of the incoming phosphorus that comes out through the outflow, and how much is lost through other processes. The addition of more phenomena, like resuspension or biological uptake, yields more complex equations. These models, not considering changes over time, are named steady state models.

Models considering changes over time are dynamic. The simplest kind is the lumped model, consisting of ordinary differential equations and with the as- sumption that the lake is perfectly mixed. The partial differential equation models, finally, add the complexity of spatial variations in 1, 2, or 3 dimen- sions.

Model building is a both complex and common task, and therefore a large

(9)

number of instructions and flow-chart diagrams exist to aid in the procedure (Jeppsson, 1996; Beck, 1983; Jørgensen, 1994; Gustafsson et al., 1982). Nor- mally the first step is selection of model type and structure. Then follows calibration, i.e. fitting of unknown model parameters so that the model out- put fit experimental data. The final step is validation, a comparison of model output to a new set of experimental data not used for the calibration. This is an important step, as it reveals the quality of the model predictions. But in many cases of ecological modelling there is such a shortage of data that it is impossible to put aside good data for the validation. Validation of ecological models is covered in an excellent way by Rykiel (1996).

Spreadsheet software is often sufficient for the steady state models. Ordinary differential equation models requires more advanced mathematical software, such as S TELLA or M ATLAB . The solution of partial differential equation models is more complex, and the software is often expensive and tailored to specific problems.

4 What’s an appropriate model for Lake Vänern?

It is generally easier to calibrate and test models based on ordinary differential equations compared to models based on partial differential equations. To use ordinary differential equations, one must assume that the lake is well mixed, and section 4.1 examines water quality data to see how well mixed Lake Vän- ern is. If the entire lake is not well mixed, but has several internally well mixed sub-basins, each sub-basin can be modelled with ordinary differential equations. Section 4.2 describes a case study concerning modelling of water exchange between different sub-basins of such a model. The study is per- formed using data from Östhammarsfjärden on the Swedish east coast, where the water is brackish and the salinity can be used as a tracer.

4.1 The spatial resolution justified by data (II)

This section analyses water quality data to determine how well mixed Lake

Vänern is to aid in the selection of model structure. It starts with vertical

differences, between surface and deep water, and then proceeds to horizontal

differences between different sub-areas in the lake.

(10)

1989 1990 1991 1992 0

1 2

Lake Number

Year

1−Jun 1−Jul 1−Aug 0

2 4

Lake Number

1989

Figure 2: The Lake Number calculated with data from location 7 (marked on the map in figure 1). When the Lake Number is over 1, the lake is stratified. When creating the closeup in the right plot, a density profile is retained for a whole month, until the next sample is taken.

It is well-known that Lake Vänern is thermally stratified in the summer. The stratification can be more or less stable. One measure of the stability is the Lake Number, presented by Imberger and Patterson (1990), and modified by Robertson and Imberger (1994) and Imberger et al. (1996). The Lake Number is the ratio of the stabilizing force of density stratification to the mixing force of the wind, and defined as:

L N = S t (H − h t )

u 2 A 3 s /2 (H − h v ) (1) where H is the maximum depth, h v is the height to the center of volume (from the lake bottom), h t is the height to the thermocline, A s is the surface area, and u is the water shear velocity due to wind calculated as:

u = 

1.56 · 10 −6 U 10 2 (2)

where U 10 is the wind speed averaged over three days 10 m above the lake surface. All quantities are in SI units, and the heights are measured from the lake bottom. S t is the Schmidt stability (m 5 s −2 ):

S t = g ρ(H)

 H

0 (h v − z) 

10 3 − ρ(z) 

A (z)dz (3)

where z is the distance above the lake bottom, g is the acceleration due to grav- ity, and ρ(z), A(z) are the water density and area at height z above the bottom respectively. The height to the center of volume (h v ) was calculated from the hypsographic curve A (z), presented in Håkanson (1978), to be 84.2 m.

When the Lake Number is above one, the lake is stratified, and when it is

below one, the lake is mixed by the wind. When the Lake Number is 1, the

(11)

0 20 40 60

Mixed

Depth (m)

13−May−1974

(a)

No thermo−

cline found 16−Jun−1974

(b)

15−Jul−1974

(c)

5 10 15 20

0 20 40

60

Depth (m)

15−Sep−1974

(d)

5 10 15 20

Mixed

Temperature ( ° C) 14−Oct−1974

(e)

5 10 15 20

Insufficient data 09−Jun−1998

(f)

Figure 3: The algorithm to find the thermocline depth through a typical year at location 1 (a-e). After the second sampling program reduction in 1995, finding the thermocline depth is hard since data is scarce (f).

wind is just sufficient to force the thermocline to the surface on the windward side of the lake, on the edge of causing upwelling phenomena (Imberger et al., 1996). At lake numbers slightly above 1 there is a tilt in the thermocline, internal seiches and partial upwelling. When the thermocline rocks back and forth with the seiche, large eddies are formed along the shores, mixing surface and deep water. This is named boundary mixing and was observed in Mono Lake, California at a Lake Number of 2 (MacIntyre et al., 1999).

A time series of the Lake Number for Lake Vänern between 1988 and 1993 is

shown in figure 2. The Lake Number is calculated from temperature profiles

measured approximately once per month at sampling location 7 (marked on

the map in figure 1). Lake Vänern is normally stratified (Lake Number above

1) for one or a few months in the summer, but as the maximum Lake Number

is about 2, the stratification is weak, and internal seiches, partial upwelling and

boundary mixing persist all through the summer. The right plot shows that

there is a lot of scatter between the sampling instants, with maximum values

exceeding 5 on calm days, but values well below 1 on windy days. For the

construction of this plot, a density profile was used for a whole month, until

the next sample was taken.

(12)

Table 1: Statistical results for vertical differences using a paired t-test.

Data Difference Confidence interval p -value Temperature

Stratified yes 6.1 ± 0.2 < 10 −6

Mixed yes 0.07 ± 0.01 < 10 −6

COD

Stratified yes 0.4 ±0.1 < 10 −6

Mixed no −0.05 ± 0.15 0 .47

Chlorophyll a

Stratified yes 0 .94 ± 0.06 < 10 −6

Mixed no −0.02 ± 0.03 0 .24

Total phosphorus

Stratified yes 0 .5 ± 0.2 < 10 −6

Mixed no −0.2 ± 0.3 0.15

Dissolved phosphorus

Stratified no −0.03 ± 0.09 0.73

Mixed no −0.1 ± 0.2 0.25

The weak stratification indicated by the Lake Number leads one to believe that the difference in water quality between the surface and bottom water is small, if it exists. A statistical evaluation was performed to find out, using a paired t-test between the average epilimnion concentration and the average hy- polimnion concentration of COD, chlorophyll a, total and dissolved phospho- rus. The depth separating the epilimnion and hypolimnion is the thermocline, the steepest temperature gradient, examined from temperature data by fitting an arctan curve. The algorithm to find the thermocline depth is demonstrated in figure 3. A ‘reference’ test for non-stratified conditions was also made, with the depth separating surface and deep water arbitrarily chosen to 20 m. Results are shown in table 1. As expected, there is no difference between surface and deep water when there is no temperature stratification. During stratified condi- tions there is a small but statistically significant difference between epilimnion and hypolimnion for COD, chlorophyll a, and total phosphorus. Therefore, the epilimnion and hypolimnion should be treated separately in a model.

The next issue concerns the horizontal differences to see if the water qual-

ity is different in different parts of the lake. The statistical test used is an

ANOVA (analysis of variance) to see if the 10 sampling locations used in

1972–1995 (marked 1–10 in figure 1) have similar water quality. If they do

not, the ANOVA is followed by a multiple comparison test, in this case the

Tukey’s honestly significant difference criterion, to see which ones are dif-

(13)

ferent. Both the ANOVA and the multiple comparison test are described in statistical textbooks, see e.g. Aczel (1999). To get a fair comparison between shallow and deep locations, the comparison is only based on surface water samples, down to 10 m depth. To remove the effects of time trends in the data, a two-way ANOVA was performed (location and time), but only the location is of interest.

The test was performed for COD, total and dissolved phosphorus, Secchi depth, and chlorophyll a. None of them indicated a fully mixed lake, and figure 4 shows the results from the multiple comparison test. Starting with COD, Dalbosjön is well mixed, and has a distinctly lower concentration than Värm- landssjön, which is also well mixed except for locations 4 and 5 near the north- ern shore. The pattern for total phosphorus is similar, but the uncertainty is larger. For Secchi depth, and chlorophyll a, there is no difference between the basins, rather large uncertainties, and the near-shore locations 2 and 4 stand out from the rest. The pattern is similar for dissolved phosphorus, except for location 2 which has moved to the main group.

For the modelling purpose both basins can be treated as well mixed horizon- tally, except for some shallow and near-shore locations. If one considers the results from COD, the basins should be modelled separately from each other, but if the analysis is based on phosphorus, Secchi depth, or chlorophyll a this is not deemed necessary.

In summary, the model to be used should have either 2 boxes, epilimnion and hypolimnion, or 4 boxes, epilimnion and hypolimnion in Dalbosjön and Värm- landssjön respectively. If near-shore areas are of interest, extra boxes have to be added to the model for each sub-area.

4.2 How to model mixing between basins (III)

With the 2 sub-basin setup, there is a need to know the amount of mixing

between the basins Värmlandssjön and Dalbosjön. As the water exchange is

not well known, a simple study was performed to find a way to model the water

exchange between interconnected basins. This study was carried out for three

basins near Östhammar on the Swedish east coast, where salt can be used as

an inert tracer. A map of the bays is shown in figure 5. Fresh water enter

through rivers into the bays, and water exchange over the narrow straits bring

brackish water from the Baltic Sea. The data available is a series of salinity

measurements from the summer of 1996, shown along with model results in

figure 6.

(14)

21 22 23 1

2 3 4 5 6 7 8 9 10

3

9 10

1 2

6 7 8

4 5

COD (mg/l)

Location

2 1

6 8 7 3 9

10 5 4 Dalbosjön

Värmlandssjö n

9 10 11

1 2 3 4 5 6 7 8 9 10

3

9 10

1 2

6 7

8

4 5

Total phosphorus ( µ g/l)

Location

2 2.5 3

1 2 3 4 5 6 7 8 9 10

3

9 10 1

2

6 7

8

4 5

Dissolved phosphorus ( µ g/l)

Location

4.2 4.4 4.6 4.8

1 2 3 4 5 6 7 8 9 10

3

9 10

1 2

6 7 8 4

5

Secchi depth (m)

Location

1.6 1.8 2 2.2

1 2 3 4 5 6 7 8 9 10

3

9 10

1

2

6 7

8

4 5

Chlorophyll a ( µ g/l)

Location

Figure 4: Multiple comparison test of 5 water quality parameters from the 10

locations that were still in use in 1995. If the uncertainty intervals

of two locations do not overlap, the water quality is significantly

different at those two locations.

(15)

18oE 12’ 24’ 36’ 48’ 19oE

9’

12’

60oN 15.00’

18’

21’

24’

18oE 12’ 24’ 36’ 48’ 19oE

9’

12’

60oN 15.00’

18’

21’

24’

10oE 15oE 20oE 25oE

54oN 57oN 60oN 63oN 66oN 69oN

Granfjärden Hunsaren

Östhammarsfjärden The Baltic Sea

Granfjärden Östhammarsfjärden

Hunsaren Q

02

Q

01

Q

21

Q

12

Q

32

Q

23

V

2

=23

·

10

6

m

3

S

2

V

1

=11

·

10

6

m

3

S

1

S

3

V

3

=35

·

10

6

m

3

Q

43

Q

34

Q

03

S

4

The Baltic Sea

Figure 5: Map of the Östhammarsfjärden study area (top), along with a schematic drawing for the double flow model (bottom). The map comes from the GSHHS database (Wessel and Smith, 1996)

The simplest model tested is the double flow model, using two continuous flows in opposite directions to approximate the real flow which oscillates back and forth at a high frequency, governed by the wind. It was also the model fi- nally used for Lake Vänern, and it is given a full description here. A schematic diagram of the model with explanation of the symbols is given in figure 5, where Q is inter-basin flow (m 3 /s), S is salinity (%), and V volume (m 3 ). As- suming that the volume of the bays is constant, dV 1 /dt = dV 2 /dt = dV 3 /dt = 0, the intermixing flows Q 12 , Q 23 , and Q 34 can be eliminated from the equations, and a salt balance model is given by:

V 1

dS 1

dt =S 0 Q 01 + S 2 Q 21 − S 1 (Q 01 + Q 21 ) (4) V 2

dS 2

dt =S 0 Q 02 + S 1 (Q 01 + Q 21 ) + S 3 Q 32 − S 2 (Q 01 + Q 02 + Q 21 + Q 32 ) (5) V 3 dS 3

dt =S 0 Q 03 + S 2 (Q 01 + Q 02 + Q 32 ) + S 4 Q 43

− S 3 (Q 01 + Q 02 + Q 03 + Q 32 + Q 43 ) (6)

where the parameters to be regressed are the flows Q 21 , Q 32 and Q 43 in order

to fit the salinities S 1 , S 2 , and S 3 to experimental data. The inlet river flows are

Q 01 = 0.25, Q 02 = 0.2 and Q 03 = 0.05 m 3 /s (Fries and Göransson, 1998). S 0

(16)

1−Jun 1−Aug 1−Oct 0.35

0.4 0.45 0.5 0.55

Salt concentration [%]

Double flow

1−Jun 1−Aug 1−Oct Wind forcing

Figure 6: Comparison of modelled (—) and measured salt concentration dur- ing the summer of 1996 for Granfjärden ( ), Östhammarsfjärden ( ), and Hunsaren (×) using the double flow and wind forcing mod- els.

is the salinity of river water (0.008 %) and S 4 is the time varying salinity of the Baltic.

Optimal values for the parameters computed using a standard unconstrained nonlinear optimiser are Q 21 = 1.6, Q 32 = 6.8 and Q 43 = 7.1 m 3 /s. The simu- lation trends given in figure 6 show results well in agreement with the experi- mental data.

With the lack of tide, the primary mixing agent in both the Baltic and in Lake

Vänern is the wind. A model for wind-driven flow across a strait developed by

Thierfelder (1995) was tested and some small modifications were introduced

and presented in paper III. As can be seen from figure 6, this model does not fit

data as well as the double flow model. It is very sensitive to strong winds, and

predicts too much mixing during storm events, such as July 9 or the storms

in early October. Therefore this model was not used for Lake Vänern, and

the model equations are not presented here. Paper III also investigated three

partial differential equation models. These models were never intended to be

used for Lake Vänern, since the previous section (4.1) found an appropriate

model structure to be 2 or 4 internally well mixed boxes.

(17)

5 Application of the L EEDS model to Lake Vänern

The demands for a water quality model stated so far is that it should model phosphorus and that it should treat surface and deep water separately, and be possible to use with one or two sub-basins. The L EEDS (Lake Ecosystem Effects Dose Sensitivity) model meets these demands. Its vertical resolution is epilimnion – hypolimnion – sediments, and apart from phosphorus it also models suspended particulate matter, which is interesting for modelling the effects of the fiber emissions from the pulp and paper industry.

L EEDS was originally a model for the phosphorus cycle, used to predict the effects of phosphorus emissions from fish farming (Håkanson and Carlsson, 1998). A model description is given by Håkanson (1999) and Håkanson et al.

(2003). The model has been merged with a model of suspended particles (SPM) (Håkanson et al., 2000), where the particulate phosphorus and the sus- pended particles are closely connected, in the way that the equations governing their behavior are similar. More detailed SPM models exist, e.g. by Weyhen- meyer et al. (1997), but this one fits the resolution of the phosphorus model and the desired model resolution for Lake Vänern. The combined model is throughout this thesis named L EEDS . A full model description including a list of symbols is given in appendix. Other present work on the model is done by Malmaeus and Håkanson (2003) and Malmaeus and Håkanson (In press).

While modifying and calibrating the model to fit Lake Vänern it was found that due to the slow dynamics of the sediment states the calibration runs need to be at least 50 years, which means it needs to start before 1972, when the water quality monitoring program started. These findings are reported in section 5.1.

Modifications to the model are described in section 5.2 and section 5.3 show the use of different model resolution, using one, two, and five sub-basins.

5.1 The need of long data series for calibration (I)

It is well known, see e.g. (Jørgensen, 1994), that one must have a long enough time series of data for calibration. It is not uncommon to see studies based on data series from just one year, e.g. Nyholm (1978) and Reynolds et al. (2000).

Longer studies exist where the nutrient loading to the lake has changed, and

one wants to model the lake response, not just a typical year. Common data

series for this type of investigations are 5–10 years, and the longest series of

(18)

nutrient data used are 23 years (Jørgensen et al., 1978) and 17 years (Varis, 1993). Therefore the calibration data available for Lake Vänern seemed to be adequate, with monthly measurements since the early 1970s.

For the first calibration, the model structure was basically unchanged from Håkanson (1999) and Håkanson et al. (2000), five parameters were fitted, and the run time was 1970–2000. But the slow dynamics of the sediment states made them dominate the whole simulation. The most important factor govern- ing the model fit to experimental data was the initial amount of x pa (phosphorus in accumulation area sediments), and not the model parameters. Starting the calibration runs in 1900 solved this problem, since the initial conditions played out their role until about 1950, and the fit was judged by the model parameters.

5.2 Model modifications for Lake Vänern (IV)

Several modifications to the L EEDS model were necessary when applying it to Lake Vänern. Three major changes were done to the model structure: the introduction of multiple basins, the modification of the equations governing the outflow of phosphorus and SPM through Göta Älv, and the inclusion of chemical oxygen demand (COD) into the model.

The results when combining the mixing model from section 4.2 with L EEDS

are given in the next section (5.3). The modifications regarding the outflow through Göta Älv and the inclusion of COD will be treated in this section, along with a short note on the calibration. Some smaller changes to the model are not mentioned here, for details see paper IV.

The old equations for the outflow were developed to handle stratified lakes, where water locked in the hypolimnion during stratified periods does not take part in the water exchange. This was accomplished with an algorithm based on the water residence time. The equations are:

D out = x de · F e · 1.386 T (

Tw+910

+0.5 ) /1.5

w

(7)

P out = x pe · F e · 1 .386 T (

Tw+910

+0.5 ) /1.5

w

(8)

S out = x se · F e · 1 .386 T (

Tw+910

+0.5 ) /1.5

w

(9)

where D out , P out , and S out are the outflow of dissolved phosphorus, particulate

phosphorus and suspended particles. x de , x pe , and x se is the amount in the epil-

(19)

1 2 3 4 5 6 7 8 10

20 30 40 50

COD (KMnO

4

consumed, mg/l)

Secchi depth (m) Data Regression

Figure 7: The regression between COD and Secchi depth.

imnion of dissolved phosphorus, particulate phosphorus and suspended parti- cles, F e (= 1) is the fraction of the inflowing water leaving the lake through the outflow (instead of evaporating), and T w is the water residence time. With the long residence time for Lake Vänern (108 months), these equations severely overestimates the outflow. They have been replaced by:

D out = θ out u Q y po

4

· 10 −9 (10) P out = θ out u Q

 y tp − y po

4

 · 10 −9 (11)

S out = θ out u Q y spm · 10 −6 (12) where u Q is the water flow through Göta Älv, and y po

4

, y tp , and y spm are the lake average concentrations of dissolved phosphorus, particulate phosphorus and suspended particles. A similar approach is suggested by Håkanson et al.

(In press). The factor θ out = 1.6, fitted during calibration, is included because the water flowing out through Göta Älv is different from the lake average.

This is probably because the bay where Göta Älv leaves (Vänersborgsviken) is different from the lake average, like many other bays are. This has not been confirmed since there are not enough measurements from Vänersborgsviken.

Though improbable, other possible explanations exist, where one is that the sampling location in River Göta Älv is badly located, and not representative of the whole river.

COD was incorporated into the model because a long data series is available

(starting in 1896), because the measurements are reliable (little noise), and

because it is a measure related to the industrial pollution from pulp and paper

mills. Using the water quality data base, a covariation was found between

COD and Secchi depth, which is already included in the model. COD was

(20)

Table 2: Parameters fitted during calibration of the L EEDS model for Lake Vänern. Old values as given by Håkanson (1999); Håkanson et al.

(2000) and values fitted in this study.

Parameter Unit Old value This study Age A months ≈ 1 · 10 5 a 696

Age et months 12 10

F r – 2 10

K spm – 1 .8 1 .7

R d month −1 2.5 · 10 −5 0

v spm m/month 42 0.63

θ out – 1 1.6

a Age A was not a constant in the original L EEDS model.

thus introduced as a linear regression from Secchi depth.

y cod = 85.9 − 14.3y sec (13) The data and regression line are shown in figure 7.

During the calibration of the model, 7 parameters were fitted. They are pre- sented in table 2. The changes are small for parameters governing the outflow ( θ out ) and primary production (K spm ). The large changes, and thus the most uncertain part of the model, is in the sedimentation (F r and v spm ) and sediment dynamics (Age A and R d ). The result of the calibration is shown in figure 8, which shows model results compared to all the reference data available. The general impression is that the model fits experimental data well before 1940 and after 1980, but the polluted conditions in the middle of the century are not modelled very well. The model results for gross sedimentation (lower left plot) is far from the reference data. This is due to the low value of the sinking speed of suspended particles (v spm = 0.63 m/month). A value of 9 instead of 0.63 gives gross sedimentation in accordance with measured data, but also gives phosphorus, Secchi depth and COD at a constant level throughout the whole simulation period, with no change in water quality during the ‘polluted’ period 1940–1980. v spm is chosen to 0.6 as a good fit for the water quality outputs is prioritized.

In summary, the L EEDS model has been modified to fit Lake Vänern by chang-

ing the equations governing the outflow of phosphorus and suspended particles

through River Göta Älv, by the introduction of chemical oxygen demand, and

by a recalibration of 7 key parameters.

(21)

LEEDS model Lake Vänern data Göta Älv data (outlet) Göta Älv (90 km downstream) 0

10 20 30

Total P (µg/l)

0 5 10

Dissolved P (µg/l)

2 3 4 5 6

Secchi depth (m)

0 20 40 60

COD (mg/l)

0 2 4 6

Chlorophyll a (µg/l)

0 0.5 1 1.5

Phytoplankton (mm3/l)

0 1 2 3 4

SPM (mg/l)

0 10 20 30 40

Dout (ton/month)

0 20 40 60

Dout+Pout (ton/month)

0 2 4 6

Sout (103 ton/month)

1 1.2 1.4 1.6 1.8

Sediment P (mg/g solids)

1900 1920 1940 1960 1980 2000

21 22 23

xsa (106 ton)

1900 1920 1940 1960 1980 2000

0 10 20 30

Sedimentation (Sg) (ton/month⋅km2)

Figure 8: Calibration results for the L EEDS model showing all reference data

available.

(22)

Dalbosjön

Värmlandssjön Kattfjorden Q

1

Q2

Q3

Qmix

Qout

Figure 9: The water flows in the 2-basin model.

5.3 Running the L EEDS model with 1, 2, and 5 sub-basins (IV)

The L EEDS model has a division of surface and bottom water, and this section shows the results of expanding it from one to two basins (representing Värm- landssjön and Dalbosjön), and further to five basins, with the inclusion of three coastal areas. All the model equations and parameters are retained from the 1- basin version, and the only modification is the addition of the mixing model, the ‘double flow’ model described in section 4.2. In the case of two basins, the flows are shown in figure 9. River inflow Q 1 and Q 2 are inputs to the model, the inter-basins flow Q 3 is a fitted parameter, and the other flows are calcu- lated with mass balances, to keep the volume of the basins constant. The best fit is achieved with Q 3 = 2300 m 3 /s, which is six times the outflow through Göta Älv. Simulation results for COD, Secchi depth and dissolved phospho- rus for both basins are shown in figure 11. The results are almost identical to the results for the one-basin model, shown in figure 8 and repeated in figure 10 to ease comparison. The difference in COD between Värmlandssjön and Dalbosjön is captured rather well by the model, but is too small to be seen in the plots.

Three coastal areas were added to the 2-basin model to construct the 5-basin

model. The first one is Kattfjorden, an enclosed bay at the northern shore

shown in figure 9. The other two are created by separating the 2 basins Värm-

landssjön and Dalbosjön into coastal and pelagial areas, with the intention to

increase the horizontal resolution of the model, and to improve the fit to exper-

(23)

imental data for the pelagial basins by using the coastal areas as buffer zones.

The fit of the 5-basin model to experimental data is poor, as seen in figure 12.

The model performs quite well for the enclosed bay Kattfjorden, but the split- ting of Värmlandssjön and Dalbosjön into near-shore and pelagial areas gives poor results for both the coastal and pelagial areas. If these coastal areas are to be included, a full recalibration of the model is needed. The 2-basin model and the Kattfjorden sub-basin did well enough with just fitting the intermixing flow. As the resolution for the main lake should be 2 basins, and there is no special interest in any of the bays, no recalibration has been performed and the 2-basin model has been used for all simulations. The lack of data for the coastal area basins, especially for the coast of Dalbosjön (figure 12(a)), also makes a recalibration difficult.

In summary, the L EEDS model works for several basins, but different param- eters may be needed in all computational elements, and it is therefore not ad- visable to have elements without reference data.

6 Comparison of eutrophication models for Lake Vänern (V)

This chapter describes a comparison of four eutrophication models applied to Lake Vänern. The comparison is focused on phosphorus because it is the lim- iting nutrient and included in all models. Two of the models are rather simple, with no horizontal or vertical differences, and designed to be part of catchment area models. One is the F YRISÅ model developed at SLU (Sonesten, In prep.) and the other one is the H BV -NP lake module developed at SMHI (Bergström, 1995; Arheimer and Brandt, 1998; Andersson et al., 2002). The other two are more complex lake models, the 2-basin version of the L EEDS model and the B IOLA model developed at SMHI (Pers, 2002; Pers and Persson, 2003).

Table 3 shows the states variables and other outputs from the four models. The

F YRISÅ and H BV -NP models only consider phosphorus and nitrogen at the

lake outflow, while the L EEDS and B IOLA models have more states and model

in-lake conditions. The most important difference between the models is the

inclusion of the sediments in the L EEDS and B IOLA models. The sediments are

needed to model long term buildup of nutrients during the polluted conditions

and the subsequent release when pollution is reduced. Another difference is the

(24)

0 20 40 60

COD (mg/l)

2 4 6

Secchi depth (m)

19000 1920 1940 1960 1980 2000

2 4 6 8

Time (year) PO43− (µg/l)

LEEDS Lake Vänern Gothenburg

Figure 10: Simulation results of the one-basin L EEDS model and experimental data for COD, Secchi depth, and dissolved phosphorus.

0 20 40 60

COD (mg/l)

2 4 6

Secchi depth (m)

19000 1920 1940 1960 1980 2000

2 4 6 8

Time (years) PO43− (µg/l)

LEEDS Dalbosjön basin Gothenburg

(a) West basin – Dalbosjön

0 20 40 60

COD (mg/l)

2 4 6

Secchi depth (m)

19000 1920 1940 1960 1980 2000

2 4 6 8

Time (year) PO43− (µg/l)

LEEDS Värmlandssjön basin Gothenburg

(b) East basin – Värmlandssjön

Figure 11: Simulation results of the two-basin L EEDS model and experimental data for COD, Secchi depth, and dissolved phosphorus.

Figure 12: (Opposing page) Simulation results of the five-basin L EEDS model

and experimental data for COD, Secchi depth, and dissolved phos-

phorus.

(25)

0 20 40 60

COD (mg/l)

2 4 6

Secchi depth (m)

19000 1920 1940 1960 1980 2000

2 4 6 8

Time (year) PO43− (µg/l)

LEEDS Coast Dalbosjön basin Gothenburg

(a) West basin, coast

0 20 40 60

COD (mg/l)

2 4 6

Secchi depth (m)

19000 1920 1940 1960 1980 2000

2 4 6 8

Time (year) PO43− (µg/l)

LEEDS Dalbosjön basin Gothenburg

(b) West basin, pelagial

0 20 40 60

COD (mg/l)

2 4 6

Secchi depth (m)

19000 1920 1940 1960 1980 2000

2 4 6 8

Time (year) PO43− (µg/l)

LEEDS

Coast Värmlandssjön basin Gothenburg

(c) East basin, coast

0 20 40 60

COD (mg/l)

2 4 6

Secchi depth (m)

19000 1920 1940 1960 1980 2000

2 4 6 8

Time (year) PO43− (µg/l)

LEEDS Värmlandssjön basin Gothenburg

(d) East basin, pelagial

0 20 40 60

COD (mg/l)

2 4 6

Secchi depth (m)

19000 1920 1940 1960 1980 2000

2 4 6 8

Time (year) PO43− (µg/l)

LEEDS Kattfjorden basin Gothenburg

(e) Kattfjorden

(26)

Table 3: The variables of the models, divided into state variables (•) and other output variables ( ◦).

Variable Model

F YRISÅ H BV -NP L EEDS B IOLA

Water temperature - - -

Lake volume - -

Dissolved phosphorus -

Particulate phosphorus - -

Total phosphorus

Phosphate in sediments - - -

Total phosphorus in sediment - - -

Organic nitrogen - - -

Dissolved inorganic nitrogen - -

Total nitrogen -

Nitrate in sediment - - -

Ammonium in sediment - - -

Oxygen - - -

Suspended particles - - -

Detritus - - -

Organic matter in sediment - -

Chemical oxygen demand (COD) - - -

Secchi depth - - -

Chlorophyll - - -

Phytoplankton - - a

Zooplankton - - -

a

In addition to phytoplankton, cyanobacteria is modelled separately.

higher physical resolution, mainly of the B IOLA model, which has 5 horizontal

‘basins’ and 34 depth layers at the deepest part of the lake. The input data needed is about the same for all models.

The model fit to experimental data for the period 1990–1995 is shown in figure 13, and the fit expressed as root mean square error for the entire calibration and validation periods is given in table 4. The fit is good, and similar for all models except for the slightly high results with B IOLA . Measured data is monthly averages from 5 locations for the Värmlandssjön basin and monthly samples for River Göta Älv. The F YRISÅ and L EEDS models aim at modelling monthly averages, while the H BV -NP and B IOLA models aim for daily values.

This gives comparison of similar variables only for H BV -NP at Göta Älv and

L EEDS at Värmlandssjön. The other comparisons are slightly wrong. The

standard deviation of sampling and analysis is 1.3 µg/l for total phosphorus

and 0.7 µg/l for dissolved phosphorus, as determined by comparing samples

(27)

0 2 4 6

Värmlandssjön basin

Dissolved P (µg/l)

BIOLA

LEEDS Measurements

River Göta Älv

LEEDS

HBV−NP

1990 1991 1992 1993 1994 1995 0

10 20 30 40

Total P (µg/l) LEEDSBIOLA

1990 1991 1992 1993 1994 1995 LEEDS, HBV−NP & FYRISÅ

Figure 13: Comparison of the model results with measured data.

Table 4: The model fit to experimental data (expressed as root mean square error). The calibration time is 1985–1992 and the validation is 1993–

2000 except for B IOLA with calibration 1990–1993 and validation 1985–1989 and 1994–2000.

Värmlandssjön Göta Älv

PO 4 TP PO 4 TP

(µg/l) (µg/l) (µg/l) (µg/l)

F YRISÅ Calibration 4 .7

Validation 3 .6

H BV -NP Calibration 1 .3 4 .7

Validation 1 .7 3 .6

L EEDS Calibration 0 .5 1 .5 1 .9 4 .8 Validation 1 .0 1 .8 1 .7 3 .7

B IOLA Calibration 2 .2 3 .1

Validation 1 .8 3 .6

from adjacent depths on the same location and sampling day.

Three scenarios were run comparing the models. The first one simulates in-

creased emissions by 40 % from the Skoghall pulp and paper mill. That means

an increase of 10 kg phosphorus per day, 150 kg nitrogen per day, and 1500

kg SPM per day. The second scenario is taken from Sonesten (In prep.) and

simulates the installation of septic tanks to all households not connected to

municipal sewage treatment. The decrease of nutrient emissions are 145 kg

phosphorus per day (14 %) and 1300 kg nitrogen per day (3 %). The third sce-

(28)

Table 5: The change in dissolved phosphorus and total phosphorus for the three scenarios. The average concentration (1990-2000) is 2 µg/l for PO 4 and 7 µg/l for total phosphorus.

Variable Model

F YRISÅ H BV -NP L EEDS B IOLA

Pulp and paper mill expansion

PO 4 (µg/l) 0 .00 0 .01 0 .01

P tot (µg/l) 0 .09 0 .12 0 .02 0 .01 Nutrient load reduction

PO 4 (µg/l) −0.6 −0.1 −0.25

P tot (µg/l) −1.3 −0.6 −0.4 −0.35

Climate change

PO 4 (µg/l) 0 −0.34 −1.2

P tot (µg/l) 0 0 −0.36 −1.2

nario simulates the effect of a temperature increase in the lake due to the green house effect. The simulation considers the years 2070–2100, and climate sim- ulations by Rummukainen et al. (2001) estimates the temperature increase in the air to be 3 degrees, which gives about 2 degrees in the water.

The changes in phosphorus predicted by the models are shown in table 5. The effects from the pulp mill expansion are small. The models predict between 0 and 1 % increase of phosphorus concentration. The effect from the installa- tion of septic tanks is much larger, as expected with a much larger change in nutrient load, up to 19 % decrease of dissolved phosphorus and 8 % decrease of total phosphorus is suggested by the models. The effects of the climate change vary a lot between the models. The F YRISÅ and H BV -NP show no effect at all. The L EEDS model shows a decrease in phosphorus about as large as for the installation of septic tanks, and B IOLA shows an even larger de- crease. Note, however, that the increased temperature is the only phenomenon included. Changed rainfall or other possible effects of the green house effect are not considered. It is worth noting that the last two scenarios are extrapola- tions outside the area of calibration and validation.

There is no clear answer to which model that is the best. They all have advan- tages and disadvantages. Starting with their description of the present situa- tion, the F YRISÅ , H BV -NP and L EEDS model are equally good, while B IOLA

has slightly poorer fit. An advantage of the F YRISÅ and H BV -NP models are

that they are part of catchment area models, and an advantage of the L EEDS

(29)

and B IOLA models that they include more variables and processes (like sed- iment storing) which makes them able to simulate more types of scenarios.

The daily input data required by H BV -NP and B IOLA is not available for river loads, and has been replaced with interpolated monthly data. B IOLA has a long run time (2 hours for a 15-year simulation on a PC), which slows down the simulations a bit. This is most troublesome during the model calibration.

The same problem exist in a smaller amount for the L EEDS model.

The conclusions from the model comparison are that the models all have sim- ilar fit to reference data, but still give rather different results when used for extrapolation in the scenarios. Finally, it is interesting to note that a tempera- ture increase of 2 degrees in the water has as large an effect on lake phosphorus as a 40 % decrease in phosphorus load.

7 Conclusions

This work examines what is a suitable model complexity for modelling nutri- ents in Lake Vänern. Statistical analyses of 30 years of water quality data show that an appropriate water quality model for Lake Vänern shall treat the epil- imnion and hypolimnion separately. Higher vertical resolution is not justified.

Horizontally, the main lake needs to be divided into 2 basins, Värmlandssjön and Dalbosjön. Shallow near shore areas, bays, and areas close to point sources need to be treated as their own basins if they are to be included in a model.

The L EEDS model, for phosphorus and suspended particulate matter, has been applied to Lake Vänern. Two major modifications were made to the model.

The first one is the modification of the terms for the outflow of phosphorus and SPM through the outlet river Göta Älv. The old regressions to handle the fact that part of the water is trapped in the hypolimnion during the summer did not work for a lake as large as Lake Vänern, and had to be replaced. The second modification is the inclusion of COD into the model.

Due to the slow dynamics of the sediment states, the calibration run for the L EEDS model needs to be at least 50 years. Otherwise the initial conditions of the sediment states will dominate the simulation and incorrect parameters will be found.

It is possible to get better horizontal resolution by using the L EEDS model for

several sub-basins within a lake and just fit the intermixing flow, as done with

the 2-basin version. Including shoreline sub-basins in the 5-basin version gave

(30)

drastically different results, and a full recalibration of the model, with different parameters in the different sub-basins, would have been necessary to achieve a good result.

The L EEDS model has been compared to three other eutrophication models applied to Lake Vänern regarding their suitability as tools for eutrophication management. The F YRISÅ , H BV -NP and L EEDS models show similar fit to the present conditions, both in calibration and validation, while the B IOLA

models shows slightly poorer fit. The two simple models (F YRISÅ and H BV -

NP , developed as parts of catchment area models) gain on their ease of use

and calibration. The more comprehensive lake models L EEDS and B IOLA gain

on their more realistic structure, and higher number of included variables.

(31)

Appendix: Description of the L EEDS model

The model description given here is identical to the one given in paper IV. An overview of the model inputs, states, outputs, and reference data is shown in figure A.1, and a flowchart diagram of the states and flow terms is shown in figure A.2. The states are further described in table A.1, and the parameters are presented in tables A.2 and A.3. A full list of symbols is given in table A.4.

Inputs States Outputs

xsa xse xsh xst xde xdh xf xpa xpe xph xpt

ychl ycod yf ypo4 ysec yspm ytp Output regressions

states

ulighturainuwind uteuth Weather

River inflow and point sources usudupuQ

Other basins usudupuQ

Chlorophyll a COD Phytoplankton Dissolved phosphorus Secchi depth SPM Total phosphorus

Data

comparison

(table 2)

Parameters (table 3)

Figure A.1: An overview of the model inputs, outputs, states, and validation data.

Table A.1: The states in the L EEDS model. All units are ton dry weight.

State Explanation

Suspended particulate matter (SPM) x sa In deep lake sediments

x se ,x sh Epilimnion, hypolimnion x st In shallow sediments

Phosphorus

x de ,x dh Dissolved in epilimnion, hypolimnion x f In phytoplankton

x pa Particulate in deep lake sediments x pe ,x ph Particulate in epilimnion, hypolimnion x pt Particulate in shallow sediments

The kernel of the dynamic model is the differential mass balances for each of the 11 states given in Table A.1. Each state is a box in figure A.2.

dx sa

dt = S sed3 − S bury (A.1)

dx sh

dt = S mix + S resh + S sed1 − S minh − S sed3 (A.2)

(32)

Table A.2: The constants used in the L EEDS model for Lake Vänern.

Symbol Value Unit Comment

Constants changed to fit the model to Lake Vänern Age A 696 month Age of A sediments a

Age et 10 month Age of ET sediments

F r 10 How much faster resuspended SPM sinks,

compared to fresh SPM from the inflows

K spm 1.7 – Calculation constant SPM

R d 0 month −1 Default diffusion rate for P v spm 0.63 m/month Default settling velocity for SPM

θ out 1.6 – How much higher the phosphorus concentra- tion in the outflow is, compared to the lake.

Constants found from maps or experimental data.

A † km 2 Lake area at normal surface level

C t pdep 0.5 mg/m 3 TP concentration in precipitation

D max † m Maximum depth

F et † – Fraction of erosion and transportation areas

Q † m 3 /s Average flow a

u rain 54 mm/month Average precipitation

u te 8 C Mean annual epilimnion temperature a

V † km 3 Lake volume a

y sec 5 m Average Secchi depth a

Constants kept at their original value from Håkanson (1999); Håkanson et al. (2000)

B d 6.32 Default bioturbation

D a 0.1 m Depth of active sediments

F pdep 0.1 – Fraction of TP in deposition being in particu- late form

F pp 0.0203 – Part of phytoplankton being phosphorus R cp 41.1 gC/gP C to P ratio in phytoplankton

R minS 0.1 month −1 Mineralization rate of SPM

R p 2.5 month −1 Phytoplankton uptake rate of phosphorus T p 0.107 month Turnover time for phosphorus in phytoplank-

ton

T r 4 C Reference temperature

† Basin specific constant. Value given in table A.3.

a

Not a constant in the original L

EEDS

model (Håkanson, 1999).

(33)

Table A.3: Values for the basin-specific constants in the different runs. See table A.2 for explanations of the symbols.

A D max Q V F et

km 2 m m 3 /month km 31 basin

Vänern 5650 106 9.6 ·10 8 153 0.50

2 basin

Dalbosjön 2068 86 9.6 ·10 8 44 0.61

Värmlandssjön 3582 106 6.5 ·10 8 109 0.43 5 basin

Dalbosjön 1718 86 9.6 ·10 8 42 0.61

Värmlandssjön 3000 106 6.5·10 8 104 0.43

Kattfjorden 54 65 1.8·10 8 1 0.19

Dalbosjön coast 350 20 3.1·10 8 4 0.90

Värmlandssjön coast 582 20 4.7·10 8 5 0.90 dx st

dt = S sed2 − S resh − S rese (A.3)

dx se

dt = S dep + u s + S die + S rese − S out − S mine − S mix − S sed1 − S sed2 (A.4) dx pa

dt = P sed3 − D diff − P bury (A.5)

dx de

dt = D dep + u d + P mine − D grow − D out − D mix (A.6) dx dh

dt = D diff + P minh + D mix (A.7)

dx pt

dt = P sed2 − P rese − P resh (A.8)

dx pe

dt = P dep + u p + P die + P rese − P mine − P mix − P out − P sed1 − P sed2 (A.9) dx ph

dt = P mix + P resh + P sed1 − P minh − P sed3 (A.10) dx f

dt = D grow − P die (A.11)

The flows of phosphorus and SPM, corresponding to the arrows in figure A.2 are described by algebraic equations.

P bury = x pa Age a

(A.12)

S bury = S sed3 (A.13)

References

Related documents

46 Konkreta exempel skulle kunna vara främjandeinsatser för affärsänglar/affärsängelnätverk, skapa arenor där aktörer från utbuds- och efterfrågesidan kan mötas eller

The increasing availability of data and attention to services has increased the understanding of the contribution of services to innovation and productivity in

This is the concluding international report of IPREG (The Innovative Policy Research for Economic Growth) The IPREG, project deals with two main issues: first the estimation of

I dag uppgår denna del av befolkningen till knappt 4 200 personer och år 2030 beräknas det finnas drygt 4 800 personer i Gällivare kommun som är 65 år eller äldre i

Detta projekt utvecklar policymixen för strategin Smart industri (Näringsdepartementet, 2016a). En av anledningarna till en stark avgränsning är att analysen bygger på djupa

Av 2012 års danska handlingsplan för Indien framgår att det finns en ambition att även ingå ett samförståndsavtal avseende högre utbildning vilket skulle främja utbildnings-,

Two pre- liminary conclusions from this ongoing project (not yet published) are that a new gravity system is needed and that 5 km resolution is sufficient for the detail

Industrial Emissions Directive, supplemented by horizontal legislation (e.g., Framework Directives on Waste and Water, Emissions Trading System, etc) and guidance on operating