02.02.2013 Views

D28: Internal seiche mixing study - Hydromod

D28: Internal seiche mixing study - Hydromod

D28: Internal seiche mixing study - Hydromod

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

Work package No. 7: Mixing by internal <strong>seiche</strong>s<br />

Lead contractor: SOG<br />

Main objective: Key Processes<br />

Strategic leader: Malgorzata Loga-Karpinska (WuT)<br />

Responsible task leader: Claude GUILBAUD (SOG)<br />

Main contributor involved: Organisation E-Mail<br />

Claude GUILBAUD SOG claude.guilbaud@sogreah.fr<br />

Eckard HOLLAN ISF isf.eurolakes@lfula.lfu.bwl.de<br />

Bernd WAHL ISF isf.eurolakes@lfula.lfu.bwl.de<br />

Kurt DUWE HYD duwe@hydromod.de<br />

Ulrich LEMMIN EPF ulrich.lemmin@epfl.ch<br />

Lars UMLAUF EPF lars.umlauf@epfl.ch<br />

Maciej FILOCHA WUT maciej.Filocha@is.pw.edu.pl<br />

Dissemination level: Public<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 1 of 92


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

Table of contents<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 2 of 92<br />

1 Introduction 4<br />

1.1 GENERAL CONSIDERATIONS 4<br />

1.2 OBJECTIVES 5<br />

2 Lac du Bourget 6<br />

2.1 INTRODUCTION 6<br />

2.2 THE ENVIRONMENT OF LAC DU BOURGET 6<br />

2.3 MEASUREMENT ON THE LAKE 7<br />

2.3.1 The data 7<br />

2.4 APPLICATIONS 11<br />

2.4.1 TELEMAC System 11<br />

2.4.2 SIMULATED SEICHES – ANALYSIS 17<br />

2.5 CONCLUSIONS 20<br />

2.6 REFERENCES 20<br />

3 <strong>Internal</strong> <strong>seiche</strong>s and <strong>mixing</strong> in Lac Léman 21<br />

3.1 THE LAC LÉMAN ENVIRONMENT 21<br />

3.2 FIELD STUDIES 22<br />

3.2.1 Modes detected 22<br />

3.2.2 Wave propagation patterns 24<br />

3.2.3 Evidence of Poincaré waves 31<br />

3.2.4 Progressive vector analysis 31<br />

3.2.5 Mixing by internal <strong>seiche</strong>s 32<br />

3.3 NUMERICAL SIMULATIONS 35<br />

3.3.1 Model Description 35<br />

3.3.2 Mixing by Long <strong>Internal</strong> Waves - Model Results 36<br />

3.4 CONCLUSIONS 42<br />

3.5 REFERENCES 44<br />

4 Maps and tables of free internal <strong>seiche</strong>s in Upper Lake Constance for practical use 46<br />

4.1 OBJECTIVE 46<br />

4.2 OUTLINE OF THE CALCULATION 47<br />

4.3 AUXILIARY FORMULATIONS FOR EVALUATION OF EIGEN-PERIODS WITH RESPECT TO DIFFERENT<br />

STRATIFICATIONS 52<br />

4.4 THE EIGEN-PERIODS INCLUDING THE CORIOLIS EFFECT 54<br />

4.5 THE HORIZONTAL STRUCTURES OF THE INTERFACE AMPLITUDES 58<br />

4.6 THE HORIZONTAL STRUCTURES OF THE CURRENTS 74<br />

4.7 REFERENCES 81


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 3 of 92<br />

5 Loch Lomond 82<br />

5.1 INTRODUCTION 82<br />

5.2 MEASUREMENT INFORMATION ON INTERNAL WAVES 84<br />

5.3 ANALYSIS AND INTERPRETATION 87<br />

5.4 REFERENCES 87<br />

6 Conclusions 91<br />

6.1 LAC DU BOURGET 91<br />

6.2 LAC LÉMAN 91<br />

6.3 LAKE CONSTANCE (BODENSEE) 92<br />

6.4 LOCH LOMOND 92


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

1 INTRODUCTION<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 4 of 92<br />

1.1 GENERAL CONSIDERATIONS<br />

During summer, heating of lake surface water leads to the constitution of a stable interface<br />

with strong thermal and density gradients at intermediate depth, the mesolimnion.<br />

Wind stress at the lake surface induces a tilting of this density interface. After the<br />

wind event, the interface oscillates in a standing wave movement (<strong>seiche</strong>s) to converge<br />

slowly to the new equilibrium position. The characteristics of these internal waves depend<br />

on the wind intensity and period between successive wind events. <strong>Internal</strong> wave<br />

amplitude may by amplified depending on the lake geometry and bathymetry. The dynamics<br />

of these phenomena are important for vertical <strong>mixing</strong> and therefore for the reaeration<br />

of mesolimnion waters<br />

Long internal waves in lakes commonly take the form of standing waves (<strong>seiche</strong>s) of frequency<br />

and form determined by basin shape and density (temperature) stratification in the<br />

water column. Recognised by Thoulet (1894) and Richter (1897), the "temperature <strong>seiche</strong>"<br />

phenomenon was first systematically explored and interpreted by Wedderburn (1912 and<br />

earlier papers). It was later shown (Mortimer 1953, 1963, 1993) to be a widespread response<br />

to thermocline tilt initiated by wind action on stratified lakes of all sizes and<br />

shapes, modified by Earth's rotation as the lake size increases. Various modes of oscillation<br />

can be excited, commonly modes No 1 and 2 (depending upon the vertical density<br />

profile) and several horizontal modes with No 1 usually dominant. The history of research<br />

on such waves has been reviewed by Mortimer (1993).<br />

In small lakes, internal <strong>seiche</strong>s are end-to-end motions along the lake axis (Lemmin<br />

1987). In large lakes, the Coriolis force transforms the end-to-end motion into rotating<br />

(amphidromic) wave patterns, if the characteristic length scale of the lake (typically the<br />

lake's width), exceeds the Rossby radius of deformation a = ci/f. Here f is the latitudedependent<br />

Coriolis parameter and ci is the speed of long internal waves in the absence of<br />

rotation. As described below, the first horizontal mode resembles a shore-hugging Kelvin<br />

wave traveling counter-clockwise (cyclonically) around the lake basin in northern hemisphere<br />

lakes, as was first demonstrated in the Lake of Geneva (Mortimer 1963) and later<br />

in Lake Biwa, Japan (Kanari 1975, 1984).<br />

For higher horizontal modes, the resulting rotating wave patterns in large lakes depend on<br />

the basin shape and size and become more complicated. Observations are often lacking.<br />

The spatial structure of higher modes can best be visualized through comparisons of numerical<br />

models with observations. In thermally stratified lakes, those observations usually<br />

take the form of vertical oscillations of near thermocline isotherms and/or current measurements.<br />

Four large and deep lakes in Europe are investigated in relation with the internal<br />

<strong>seiche</strong>s phenomenon: Lac du Bourget, Lake Constance, Lac Léman and Loch Lomond.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 5 of 92<br />

Different approach were used:<br />

• Lac du Bourget : Analysis of measurements available and numerical simulation<br />

with a three-dimensional model of the internals waves<br />

• Lake Constance : Estimation of the internal waves through linear theory model<br />

• Lac Léman: Analysis of measurements and numerical simulation with a threedimensional<br />

model of the internals waves<br />

• Loch Lomond: Measurements analysis<br />

1.2 OBJECTIVES<br />

To quantified the influence of internal <strong>seiche</strong>s it is crucial to know the properties of possible<br />

internal gravity waves (periods, amplitude, located depth, Coriolis influence, associated<br />

current characteristics).Then the influence of internal <strong>seiche</strong>s on <strong>mixing</strong> processes<br />

and there impact concerning the water management, constructions or executives<br />

measures could be investigated.<br />

During the period of summer stratification, <strong>mixing</strong> in lakes can be cause by vertical<br />

<strong>mixing</strong>, boundary <strong>mixing</strong>, interbasin density current and river inflows. In Lac Léman,<br />

river inflows and interbasin density currents are of little importance during summer<br />

stratification. Vertical <strong>mixing</strong> can be related to wind induced shear or to internal <strong>seiche</strong><br />

and internal wave activity.<br />

The assessment of the local intensity of internal <strong>seiche</strong>s due to their variable structure<br />

is of considerable concern in applications of water management, certain water constructions<br />

and other executive measures, for which essential information on transient<br />

currents and corresponding water displacements are required.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

2 LAC DU BOURGET<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 6 of 92<br />

2.1 INTRODUCTION<br />

The Lac du Bourget is the smallest and the less deep of the studied lakes of the project.<br />

There is very few data available in the Lac du Bourget in relation to the internal gravity<br />

waves (P.E. Bournet 1996). These data were analysed in order to identify and quantify<br />

the <strong>seiche</strong>s in the lake. Three episodes located principally in winter have been selected.<br />

The numerical model (TELEMAC-3D) was used to reproduce the ”real world”. We tried<br />

to reproduce the hydraulic respond of the stratified lake under the wind forcing. The<br />

comparison between the model results and the measurement is shown general qualitative<br />

agreement.<br />

2.2 THE ENVIRONMENT OF LAC DU BOURGET<br />

The Lac du Bourget is the largest natural<br />

lake in France located in the French Alps<br />

(area: 44.5 km 2 , length: 18 km, average<br />

depth: 85 m; maximum depth: 145 m,<br />

lake surface altitude: 231 m). Lac du<br />

Bourget is set in a depression that geologists<br />

call a syncline, resulting from the<br />

folding of the Alpine chain during the Tertiary<br />

era. The depression was further<br />

B point<br />

deepened by almost 145 m.<br />

Figure 2 : Lac du Bourget It is oriented in<br />

the North-south direction and it is composed<br />

of two main basins with similar<br />

size: the northern basin with the maximum<br />

depth (145 m); and the Southern<br />

one with a maximum depth of 70 m.<br />

The two basins are separate by a con-<br />

River Sierroz traction of the coastline. The maximum<br />

width is about 3.2 km at Grésine and the<br />

South basin<br />

minimum width at Saint-Innocent is 1.6<br />

T point<br />

km. The rivers Leysse and Sierroz are<br />

the two main tributaries recharging the<br />

lake. At the northern end, near the<br />

Chautagne marshes, Savière canal links<br />

the lake with the river Rhone, is the main<br />

outflow.<br />

Canal de Savière<br />

North basin<br />

Figure 1 :<br />

Lac du Bourget<br />

River Leysse


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 7 of 92<br />

2.3 MEASUREMENT ON THE LAKE<br />

P.E. Bournet (1996) made the major measurements project on the Lac du Bourget between<br />

1994 and 1996. During this period 18 campaigns of vertical temperature profiles<br />

at the T point were conducted. The thermistor chain recorded temperature every 10 min<br />

over 9 spaced depths between –10 m to –51m. By a spectral analysis of the measurements<br />

P.E Bournet extracted the energy density as a function of the depth.<br />

From this analysis we have extracted three periods to define scenarios that are interesting<br />

to be reproduced by the numerical model. In the following table, we have compiled<br />

for each scenario the first horizontal mode period and the more energetic depth.<br />

Scenario Dates 1er Mode<br />

Period<br />

1 12-19 December<br />

1995<br />

2 03-09 April 1994<br />

09-15 April 1994<br />

3 27/Nov.<br />

1994<br />

to 7/Dec.<br />

Depth<br />

71 h –50 m<br />

50 h<br />

85 h<br />

43 h -25 m<br />

-20 m to –30 m<br />

-40 m<br />

2.3.1 The data<br />

In the Figure 3, Figure 5 and Figure 7 we present the temperature time series measured<br />

at the T point (see Figure 2) for the chosen scenarios at different depth. The<br />

curves are shifted for comprehension.<br />

For the scenario 1 (Figure 3), The decrease in time of the temperature at the depth –<br />

23.0m (similar for the other depth) is due to refreshment of the air above the lac. The existence<br />

internal waves at depth –35.0 and –41.0m are directly identified. This <strong>seiche</strong> is<br />

related to the high wind speed at the surface of the lac on the 13 and 14 September<br />

(see Figure 4).<br />

For the scenario 2 (Figure 5) the internal waves are not well identified, particularly due<br />

to the meteorological variability (see Figure 6).<br />

The measurements for the scenario 3 (Figure 7), shows a well-developed internal wave<br />

located around –23.0 m but without correlation with the wind speed!


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

8.81°C<br />

7.93°C<br />

7.34°C<br />

12.12.95 14.12.95 16.12.95 18.12.95 20.12.95<br />

4<br />

22.12.95<br />

Figure 3 : Temperature series at the T point; Scenario 1. Curves are shifted by 1°C.<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

Figure 4 : Wind speed series at Voglans meteorological station; Scenario 1.<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Wind Speed<br />

-23.0 m<br />

-35.0 m<br />

-41.0 m<br />

12/12/95 13/12/95 14/12/95 15/12/95 16/12/95 17/12/95 18/12/95 19/12/95 20/12/95<br />

9<br />

8,5<br />

8<br />

7,5<br />

7<br />

6,5<br />

6<br />

5,5<br />

5<br />

4,5<br />

Page 8 of 92


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

8.00°C<br />

6.39°C<br />

6.09°C<br />

3<br />

03/04/94 04/04/94 05/04/94 06/04/94 07/04/94 08/04/94 09/04/94 10/04/94 11/04/94 12/04/94<br />

Figure 5 : Temperature series at the T point; Scenario 2. Curves are shifted by 1°C.<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

Figure 6 : Wind speed series at Voglans meteorological station; Scenario 2.<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Wind Speed<br />

-10.0 m<br />

-23.0 m<br />

-29.0 m<br />

0<br />

03/04/94 05/04/94 07/04/94 09/04/94 11/04/94 13/04/94 15/04/94<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

Page 9 of 92


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

8.19°C<br />

7.48°C<br />

7.21°C<br />

30/11/94 01/12/94 02/12/94 03/12/94 04/12/94 05/12/94 06/12/94 07/12/94<br />

Figure 7 : Temperature series at the T point; Scenario 2. Curves are shifted by 1°C.<br />

Figure 8 : Wind speed series at Voglans meteorological station; Scenario 3.<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

-23 m<br />

-26 m<br />

-29 m<br />

Wind Speed<br />

0<br />

28/11/94 30/11/94 02/12/94 04/12/94 06/12/94 08/12/94<br />

11<br />

10.5<br />

10<br />

9.5<br />

9<br />

8.5<br />

8<br />

7.5<br />

7<br />

6.5<br />

Page 10 of 92


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

2.4 APPLICATIONS<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 11 of 92<br />

2.4.1 TELEMAC System<br />

The TELEMAC system is a powerful integrated modelling tool for use in the field of<br />

free-surface flows (J.M. JANIN 1992).<br />

The TELEMAC-3D software solves 3D hydraulic equations (with the assumption of hydrostatic<br />

pressure conditions and time-dependent surface) and transport-diffusion<br />

equations for intrinsic values (temperature, salinity, concentration). The main results<br />

obtained at each point of the computational mesh are velocity in three directions and<br />

the concentration of transported quantities. The main result for the surface mesh is the<br />

water depth.<br />

The space is discretised in the form of an unstructured grid of triangular elements<br />

(Figure 9 on the left), which means that it can be refined particularly in areas of special<br />

interest.<br />

Bathymé-<br />

Figure 9 : Mesh and bathymetry used by TELEMAC-3D


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 12 of 92<br />

2.4.1.1 TELEMAC-3D Equations<br />

The code solves the three-dimensional hydrodynamic equations under the following assumptions:<br />

• Navier-Stokes 3D equations with free surface changing in time,<br />

• Negligible density variation in the mass conservation equation,<br />

• Hydrostatic pressure assumed,<br />

• Boussinesq approximation for momentum.<br />

Given these assumptions, the following 3D equations are solved:<br />

∂u<br />

∂u<br />

∂u<br />

∂u<br />

1 ∂p<br />

∂ � ∂u<br />

� ∂ � ∂u<br />

� ∂ � ∂u<br />

�<br />

+ u + v + w = − + �νH<br />

� +<br />

F<br />

t x y z x x x y<br />

�<br />

�νH<br />

y<br />

�<br />

� + �νH<br />

� +<br />

∂ ∂ ∂ ∂ ρ0<br />

∂ ∂ � ∂ � ∂ � ∂ � ∂z<br />

� ∂z<br />

�<br />

∂v<br />

∂v<br />

∂v<br />

∂v<br />

1 ∂p<br />

∂ � ∂v<br />

� ∂ � ∂v<br />

� ∂ � ∂v<br />

�<br />

+ u + v + w = − + �νH<br />

� + H<br />

H F<br />

t x y z 0 y x x y �<br />

�ν<br />

y �<br />

� + �ν<br />

� +<br />

∂ ∂ ∂ ∂ ρ ∂ ∂ � ∂ � ∂ � ∂ � ∂z<br />

� ∂z<br />

�<br />

S ∆ρ<br />

p = ρ0g(<br />

S − z)<br />

+ ρ0g�<br />

dz<br />

z ρ0<br />

∂u<br />

∂v<br />

∂w<br />

+ + = 0<br />

∂x<br />

∂y<br />

∂z<br />

∂T<br />

∂T<br />

∂T<br />

∂T<br />

∂ � ∂T<br />

� ∂ � ∂T<br />

� ∂ � ∂T<br />

�<br />

+ u + v + w = �νHT<br />

� + HT + �νHT<br />

� + Q<br />

t x y z x x y �<br />

�ν<br />

y �<br />

�<br />

∂ ∂ ∂ ∂ ∂ � ∂ � ∂ � ∂ � ∂z<br />

� ∂z<br />

�<br />

with:<br />

h (m) water depth Zf (m) bottom elevation<br />

S (m) free surface elevation<br />

u, v, w<br />

(m/s)<br />

ρ 0 (X) reference density<br />

velocity components ∆ρ (X) variation in density<br />

T (°C) active or passive<br />

tracer<br />

t (s) time<br />

P (X) pressure x, y (m) horizontal space<br />

components<br />

g (m/s2)<br />

νH,νz<br />

(m2/s)<br />

νHT,νzT<br />

(m2/s)<br />

acceleration due to<br />

gravity<br />

velocity diffusion<br />

coefficients<br />

tracer diffusion coefficients<br />

Fx, Fy<br />

(m/s2)<br />

Q (tracer<br />

unit)<br />

source terms<br />

tracer source or<br />

sink<br />

In order to reproduce properly the internal waves with TELEMAC-3D, we need horizontal<br />

level meshes between the surface down to the thermocline. So we cut the<br />

bathymetry above the depth –70.0m bellow the surface. Only the bathymetry below this<br />

depth is the real one. The resulted bathymetry is presented on Figure 9 (right).<br />

x<br />

y


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 13 of 92<br />

2.4.1.2 Lac du Bourget application<br />

To compare the results from the model with the measurement at the T point, we simulate<br />

and analyse the scenario 1.<br />

The simulation characteristics for the model run are:<br />

• initial lac status without velocity,<br />

• horizontal homogeneous stratification, initial vertical profile of temperature from<br />

the measurements (Figure 10),<br />

• outflow and inflow neglected.<br />

0<br />

6<br />

-20<br />

6.5 7 7.5 8 8.5 9 9.5<br />

-40<br />

-60<br />

-80<br />

-100<br />

-120<br />

-140<br />

Initial temperature<br />

Figure 10 : Initial temperature of the lac.<br />

The measured wind and temperature at the Voglans station (between 12 of december<br />

to 22 of december 1995), are impose all over the lac. The time series of the lac surface<br />

boundary condition are reported on the Figure 11.<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

Int ensit é du vent<br />

Températ ure<br />

0<br />

-2<br />

12/12/95 14/12/95 16/12/95 18/12/95 20/12/95 22/12/95<br />

Figure 11 : Vent et Température de l’air à Voglans.<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 14 of 92<br />

On the Figure 12, we have reported the time evolution of the lac temperature at the<br />

depth –23.0m for the measurement and the simulation. The decreasing of the water<br />

temperature during the simulated period is connected the a lower air temperature compare<br />

to the lac temperature associated with high wind intensity (Figure 11) which mixed<br />

the upper water level of the lac.<br />

8.9<br />

8.7<br />

8.5<br />

8.3<br />

8.1<br />

7.9<br />

7.7<br />

7.5<br />

Profondeur : -23.0 m<br />

Simulation pt T<br />

Mesures<br />

Simulation pt B<br />

12/12/95 13/12/95 14/12/95 15/12/95 16/12/95 17/12/95 18/12/95 19/12/95 20/12/95 21/12/95 22/12/95<br />

Figure 12 : Temporal evolution of the temperature at –23.0 m below the surface. Comparison between<br />

the measurement and the results from the simulation at the T point. Results from the simulation<br />

at the B point<br />

The comparison at the T point (south part of the lac see Fehler! Verweisquelle konnte<br />

nicht gefunden werden.) is showing a good agreement, until the 18 of December,<br />

between the measurement and the simulated results (purple curves and dark blue<br />

curve). After this date the shift (about 0.3°C) is constant between the two curves. The<br />

rapid variation of the temperature measured (12/12, 16/12 and 18/12) is not reproduce<br />

by the model. The rapid decreasing of temperature the 16/12 is reflecting the <strong>seiche</strong>s<br />

existed below.<br />

The comparison between the simulation results at the T point and the B point demonstrate<br />

that at the considered depth the wind action is homogeneous all over the lake<br />

between the surface and –21.0 m (see Figure 10).<br />

At the depth –35.0 m (Figure 13), the measurement at the T point (curve in purple)<br />

show clearly the existence of the internal gravity wave at this depth. The amplitude of<br />

the temperature is bigger than 1°C.<br />

At this depth, the results of the simulation are in a good agreement until mid of the<br />

15/12. After this date, the first oscillation of the temperature is well reproducing in time<br />

but the amplitude is two times smaller than the measurements. At the end of the simulated<br />

time the shift in time and amplitude is bigger.<br />

The amplitude of the <strong>seiche</strong>s at the B point is lower than at the T point, and there is a<br />

shift in phase.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 15 of 92<br />

The development of the <strong>seiche</strong> is complete at the depth –41.0 (Figure 14), the measurements<br />

at the T point (curve in purple) illustrate this point. The amplitude of the temperature<br />

is bigger than 2°C and decrease with time.<br />

At this depth, the results of the simulation are in a very good agreement until mid of the<br />

15/12. After this date, the next oscillation of the temperature is relatively well reproducing<br />

in time but the amplitude is underestimated compared to the measurements. At the<br />

end of the simulated time the shift in time and amplitude is bigger.<br />

9<br />

8.5<br />

8<br />

7.5<br />

7<br />

6.5<br />

Profondeur: -35.0 m<br />

Simulation pt T<br />

Mesures<br />

Simulation pt B<br />

12/12/95 13/12/95 14/12/95 15/12/95 16/12/95 17/12/95 18/12/95 19/12/95 20/12/95 21/12/95 22/12/95<br />

Figure 13 : Temporal evolution of the temperature at –35.0 m below the surface. Comparison between<br />

the measurement and the results from the simulation at the T point. Results from the simulation<br />

at the B point<br />

8.5<br />

8.3<br />

8.1<br />

7.9<br />

7.7<br />

7.5<br />

7.3<br />

7.1<br />

6.9<br />

6.7<br />

6.5<br />

Profondeur : -41.0 m<br />

Simulation pt T<br />

Mesures<br />

Simulation pt B<br />

12/12/95 13/12/95 14/12/95 15/12/95 16/12/95 17/12/95 18/12/95 19/12/95 20/12/95 21/12/95 22/12/95<br />

Figure 14 : Temporal evolution of the temperature at –41.0 m below the surface. Comparison between<br />

the measurement and the results from the simulation at the T point. Results from the simulation<br />

at the B point


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 16 of 92<br />

2.4.1.3 Frequency analysis<br />

A simple decomposition of the <strong>seiche</strong>s is used through a trigonometric function decomposition<br />

of the calculated temperature of the lac at the T point at –41.0 min based on<br />

the following formula:<br />

� i<br />

( ω t + )<br />

−αi<br />

S = C + A e cos ϕ<br />

i<br />

Where S is the simulated temperature, C a constant and i is the order of the decomposition.<br />

The first level of decomposition gave the following parameters:<br />

C = 7.<br />

6<br />

A1 = −0.<br />

9<br />

1 10800 = ϕ<br />

267600 − t<br />

α1<br />

=<br />

500000<br />

2π<br />

ω 1 =<br />

300000<br />

This gave a period close to 83 hours for the first mode. On Figure 15 is shown the<br />

graphical comparison between the simulated temperature at T point at –41.0m and the<br />

first level of trigonometric decomposition. For the selected period (between 200000s<br />

and 700000s) the both curve are very close.<br />

The second level of the decomposition gave the following parameters:<br />

2 0.<br />

12 = A<br />

2π<br />

ω 1 =<br />

94000<br />

i<br />

1 = ϕ<br />

i<br />

7200<br />

With this method the second mode is estimated to 26 hours.<br />

8.5<br />

8.3<br />

8.1<br />

7.9<br />

7.7<br />

7.5<br />

7.3<br />

7.1<br />

6.9<br />

200000 300000 400000 500000 600000 700000<br />

Figure 15: Simulated temperature at T point at –41.0m (S in dark Blue), first level of trigonometric<br />

decomposition (in red), first plus second level in light blue.<br />

S<br />

First level<br />

1+2


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

2.4.2 SIMULATED SEICHES – ANALYSIS<br />

2.4.2.1 Rotation<br />

The Figure 16 shows the temperature<br />

iso-contour level 7.6°C at six different<br />

times at –41.0 m depth. These times<br />

are chosen in order to illustrate the rotation<br />

of the internal wave around the<br />

amphidromic point due to the coriolis<br />

forcing. The coloured stars are associated<br />

to the position of the temperature<br />

iso-contour that has the same colour.<br />

The black arrow that follow the stars<br />

explicit the wave rotation direction.<br />

The both lines Red and Purple, which<br />

are superimposed, are related to the<br />

wave period of the <strong>seiche</strong>s (80 hours).<br />

The intersection of the six curve is located<br />

at the middle of the lac, determine<br />

the amphidromic point of the internal<br />

wave.<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 17 of 92<br />

Figure 16 : Coriolis effect on the internal wave propagation.<br />

°C<br />

°C<br />

°C<br />

°C<br />

°C<br />

°C


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 18 of 92<br />

2.4.2.2 Vertical displacement<br />

In order to evaluate the amplitude of the wave and the affected area we show on Figure<br />

17 (depth –35.0 m at left and –41.0 m on the right), the difference between the temporal<br />

(during all the simulation) maximum of the temperature and the initial temperature at<br />

the considered depth.<br />

The maximum difference is obtain on the depth –41.0 m, as the analysis made by P.E.<br />

Bournet shown. At the middle of the lac, the difference is zero in both cases, that confirm<br />

the amphidromic point (Figure 16). The maximum value are located on the west<br />

side of the lac. The North basin is more exposed to high amplitude of internal wave.<br />

Figure 17 :Difference between the maximum temperature and the initial one (-35.0 m on the right, -<br />

41.0 m on the left)


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 19 of 92<br />

2.4.2.3 Vertical velocity<br />

The maximum of the vertical velocity, during all the simulation time, for the depth –<br />

35.0m and –41.0m are shown on Figure 18.<br />

We observe a very similar value for both levels. The maximum of the vertical velocity is<br />

concentrated along the coast line with some spot located at particular points.<br />

Figure 18 : Maximum vertical velocity during the simulation (-35.0 m on the right, -41.0 m on the<br />

left)


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 20 of 92<br />

2.5 CONCLUSIONS<br />

The analysed measurements are shown the existence of internal gravity wave in the<br />

Lac du Bourget with a period between 40hours to 80 hours depending on the vertical<br />

stratification structure.<br />

The simulation using the TELEMAC-3D model is used to reproduce one <strong>seiche</strong> event<br />

on the Lac du Bourget. The numerical results are in a quite good qualitative agreement<br />

with the measurements.<br />

The numerical results are given some interesting new information about the hydraulic<br />

respond of the lake under wind forcing. After the generation of the internal <strong>seiche</strong>s, the<br />

wave propagation is showing a rotating structure, due to the Coriolis influence, with an<br />

amphidromic point at the middle of the lake (Figure 16).<br />

In spite of the bad representation of the bathymetry along the coast, the maximum dissipation<br />

of the internal wave illustrated by the two figures (Figure 17 and Figure 18) is<br />

located all around the coastline of the lake. And the maximum vertical velocity is less<br />

than 1. 10 -3 m/s.<br />

2.6 REFERENCES<br />

Bournet PE. 1996. Contribution à l’étude hydrodynamique et thermique du Lac du<br />

Bourget. PhD Thesis.<br />

Janin JM, Lepeintre F, and Pechon P. 1992. TELEMAC-3D: a finite element code to<br />

solve 3D free surface flows problems. Proceedings of computer modelling of seas and<br />

Coastal regions. Southampton, UK.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

3 INTERNAL SEICHES AND MIXING IN LAC LÉMAN<br />

3.1 THE LAC LÉMAN ENVIRONMENT<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 21 of 92<br />

The Lac Léman (Figure 19) is a warm monomictic lake situated between Switzerland<br />

and France. It is curved in shape and composed of two main basins: a deep central<br />

eastern basin (310 m maximum depth, 157 m mean depth, mean width 10 km) called<br />

Grand Lac ("big lake") and a relatively small and narrow section in the west called Petit<br />

Lac ("small lake," maximum depth 70 m; mean width around 4 km). The lake has a<br />

total length of 70.2 km and a width of 13.8 km in the central part which corresponds to<br />

2.4 Rossby radii under typical summer stratification conditions. The eastern part of the<br />

lake is surrounded by high mountains sheltering it from most strong winds. The central<br />

and western part of the lake form part of the Swiss central plateau. The windfield over<br />

the lake is affected by the mountains and the plateau. It is consequently dominated by<br />

events of strong winds from the northeast and southwest which may last from several<br />

hours to several days. The winds from the northeast have been observed to cause<br />

thermocline depressions of more than 20 m in the Petit Lac (Lemmin and D'Adamo<br />

1996). These wind events can be considered as the principal forcing to initiate internal<br />

<strong>seiche</strong>s.<br />

�<br />

����<br />

����<br />

��<br />

��<br />

������<br />

���������<br />

��<br />

�<br />

���<br />

��<br />

���<br />

����<br />

������<br />

��������<br />

Figure 19 : Map of the Lac Léman (Lake Geneva) with surrounding topography and positions of<br />

recording stations discussed in this text. Depth contours are given with reference to the water<br />

surface level which is at an elevation of 371 m above sea level.<br />

��<br />

��<br />

�<br />

���<br />

�����<br />

�� ��


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

3.2 FIELD STUDIES<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 22 of 92<br />

Field measurements concerning temperature and currents have been carried out in the<br />

Lac Léman over the last twenty years using moored instruments. The recording interval<br />

was 30 min. or 60 min. and instruments were deployed in campaigns lasting for several<br />

months. Most of these data were collected along the northern shore of the lake at<br />

mooring stations covering the slope down to about 175 m during fall stratification. However,<br />

some measurements were also made during summer periods. The data were<br />

analyzed for the presence of internal <strong>seiche</strong> modes using spectral analysis. For spectral<br />

analysis of the records, a standard Fast Fourier Transform with segment overlap was<br />

used.<br />

3.2.1 Modes detected<br />

Modes were identified by comparing the periods P corresponding to prominent spectral<br />

peaks with those predicted for a two-layered approximation fitted to the lake dimensions<br />

and density distribution using the Merian formula<br />

P n =2L b<br />

(h1 +h2) nh1h2 g(ρ2 − ρ1) ρ2 where Lb is the basin length, h1 and h2 are the depth of the epilimnon and the hypolimnion<br />

and ρ1 and ρ2 are the respective densities. n is the mode number. In the calculations<br />

we assumed, for the summer (values for the fall in parenthesis) a top layer h1 =<br />

15 (25) m; a bottom layer h2= 175 (165) m; a top layer temperature T1 = 19 (8) °C; and<br />

a bottom layer temperature T2= 5.5 °C. The analysis of all the data indicates that, independent<br />

of season and station location, only certain modes are excited. Differences<br />

were found between the number and type of modes observed at different locations in<br />

the lake in particular with distance from the shore.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 23 of 92<br />

Figure 20 : Cumulative energy spectra of alongshore current component at stations P1 and P2 (for<br />

positions see Figure 19). The maximum at the nearshore station P1 is at the Kelvin wave period<br />

while that for the offshore station P2 is at the Poincaré period. (cpd = cycles per day; modified<br />

from Bohle-Carbonell, 1986)<br />

At the stations closest to shore, the most prominent <strong>seiche</strong> is the first mode (n = 1;<br />

called L1 hereinafter; Figure 20). Its period is near 81.5 h in summer, increasing to 131<br />

h in fall, as the density structure of the water column changes and ci decreases. In the<br />

spectra from the narrow western end of the lake (not shown here) the L1 mode response<br />

is always seen most clearly. Recently, timeseries measurements were carried<br />

out during summer stratification with a current meter placed about 5 m above the lake<br />

bottom on the southern side of the deep central plateau at station S (Figure 19). Spectra<br />

from this station (Figure 21) show again a broad peak at the Kelvin <strong>seiche</strong> period.<br />

Seiche modes n = 2 to 9 were not observed in any spectra.<br />

The first cross basin or transverse mode (called T1 hereinafter), with summer and fall<br />

periods of 10.7 and 13.5 hours respectively. It is seen most clearly in the spectra at offshore<br />

stations in the central part of the lake basin (Figure 20) and is often found in the<br />

eastern part. Modes higher than 10 cannot be detected with certainty because of the<br />

cut-off imposed by the time step of the data records. The T1 period is also clearly seen<br />

in spectra from the recent deep measurements at S (Figure 21).


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

10 3<br />

Lake Geneva, midlake station, depth 310m, from 6 July 14:00 to 12 October 24:00, 2001<br />

10 2<br />

10 1<br />

10 0<br />

10 -7 10 -6 10 -5 10 -4 10 -3<br />

10 -1<br />

north component at 304m<br />

east component at 304m<br />

frequency (cph)<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 24 of 92<br />

Figure 21 : Spectra of north and east component of the currents at 304m depth recorded during<br />

summer 2001 at station S. For station location see Figure 19.<br />

3.2.2 Wave propagation patterns<br />

3.2.2.1 Mode L1<br />

It should be noted that the thermocline oscillations are accompanied by oscillations in<br />

lake surface level which are out-of-phase with and about 1/1000 smaller in amplitude<br />

than the oscillations of thermocline isotherms. It is therefore possible (e.g. Caloi et al.<br />

1961; Sirkes 1987) to use numerically filtered deviations of surface level from equilibrium<br />

to follow the progress of internal motions. This possibility arose for the Lake of<br />

Geneva, because the Swiss Service Fédérale des Eaux, (SFE, 1954) published tables<br />

of water levels measured in 1950 by 13 high-precision water level recorders spaced<br />

around the lake shore (see Fig. 1). We have analyzed these limnigraph records for the<br />

wave propagation pattern. Cross spectral analysis reveals the wave propagation pattern.<br />

Coherence and phase were determined between pairs of stations.<br />

Coherence is always high at the mode L1 period and falls off rapidly below the confidence<br />

limit for surrounding periods. A calculation was made for all possible combinations.<br />

The progression for the L1 mode during the summer period is presented in Figure<br />

22. In that case, station 2 was taken as the reference station and pairs were formed<br />

with all other stations around the lake basin. Indicated for each station are coherence


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 25 of 92<br />

and phase angle. Included, for comparison, is the L1 mode amphidromic pattern obtained<br />

from the numerical model under the conditions specified above.<br />

During summer, coherence was high for most station pairs. For the stations along the<br />

northern shore, coherence increases again from E to W and became high for those<br />

stations in the narrow Petit Lac. The phase angles calculated from the data can be<br />

compared to those predicted by the numerical model. Along the southern shore, satisfactory<br />

agreement in phase angle was found; and cyclonic progression was clearly established.<br />

Agreement was less satisfactory on the northern shore, particularly, at the<br />

entrance to the Petit Lac at stations 10 and 11, even though the coherence was high.<br />

At the western end of the lake, agreement was again good, indicating that the L1 <strong>seiche</strong><br />

mode completed a full cycle around the lake. During fall, interstation coherence decreased<br />

from W to E along the southern shore.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 26 of 92<br />

Figure 22: The L1 longitudinal mode during the summer interval 4 June to 24 Aug. 1950: top panel,<br />

structure as predicted by the model. At each limnigraph station marked by an encircled number in<br />

the top panel, phase and coherence with reference to station 2 obtained from the analysis of the<br />

waterlevel data are indicated. Middle panels, spectra of SFE water level fluctuations at station 2<br />

and 7. The small peak at the L3 mode period is not significant; T1 is the first transversal mode<br />

wave to be discussed below. Bottom panels, coherence and phase angle between 2 and 7 (out of<br />

phase at -180°). Coherence and phase at other stations relative to 2 are shown in the top panel.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

a.1 Evidence for "Kelvin-<strong>seiche</strong>" response<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 27 of 92<br />

Initial analysis of part of the water level data had already shown (Mortimer 1963) that<br />

periodicity in the surface elevation signals corresponded to the L1 mode "Kelvin<br />

<strong>seiche</strong>." That signal was visually correlated with the thermocline oscillation at Geneva.<br />

The surface elevation signal progressed around the lake perimeter at a nearly constant<br />

speed corresponding to a L1 mode internal <strong>seiche</strong>.<br />

The results of the present statistical analysis indicate that, for the L1 mode, coherence<br />

is high in the narrow western basin falling off toward the eastern end of the lake where<br />

the amplitude is greatly reduced and the wave form is perturbed by either T1 mode<br />

waves or non-wave effects.<br />

Analysis of lake temperature and current data has previously shown (Bohle-Carbonell<br />

1986) that typically one or two whole-basin L1 circuits are completed when post-storm<br />

calm persisted for long enough. Evidence for the Kelvin wave character of the L1 mode<br />

wave also came from current and temperature recordings, carried out in the western<br />

part of the Grand Lac in 1982/1983 (Mortimer 1993) from December to January when<br />

the lake was still weakly stratified with a thermocline at about 90 m depth. Parts of this<br />

analysis are reproduced in Figure 23. Station A1 is close to shore while station A3 is<br />

further offshore (for station location see Figure 19). Three wind events occurred between<br />

9 and 20 December.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 28 of 92<br />

Figure 23: Current and temperature records of fall/winter 1982/1982. Winds were measured at<br />

Cointrin Airport near Geneva at the west end of the lake basin. The wind speed squared is multiplied<br />

with the direction of the wind. Positive winds are coming from the NE; negative winds are<br />

coming from the SW (adapted from Mortimer, 1993). Measurement depths are indicated on each<br />

panel. For station locations, see Figure 19.<br />

At roughly 100 h after the start of each wind event, the currents at 15 m at mooring A1<br />

reversed suddenly from eastgoing to westgoing at the mooring nearest shore. Those<br />

reversals, R1, R2, R3, and R7, were each accompanied by a sudden depression of the<br />

thermocline (i.e. a sudden temperature rise) at mooring A1, marking the passage of a<br />

surge. Similar saw-toothed "waves" were also seen at moorings further offshore at A3<br />

but with reduced amplitude. The reversals in current direction, however, were confined<br />

to the nearshore instrument, A1.Figure 23 indicates a pattern, which starting on the<br />

south shore, had traveled around the lake (the Petit Lac was destratified and took no<br />

part in the <strong>seiche</strong>) at internal <strong>seiche</strong> speed.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 29 of 92<br />

After the wind event R3, as would be expected for a Kelvin wave, a free oscillation persisted<br />

for three cycles with amplitudes of the current reversal R3 to R5 and amplitudes<br />

of the associated temperature waves decreasing steadily in time. It can be seen that<br />

each time a strong current reversal in the upper layer occurred at station A1, but not at<br />

station A3, in accordance with the Kelvin wave exponential decay of wave amplitude<br />

with distance from shore. The temperature records again supported the Kelvin wave<br />

pattern: pronounced signals were observed at A1; at A3 (not shown) the amplitude was<br />

rather small.<br />

At that time of the year, the thermocline was below the depth of the Petit Lac basin.<br />

The path of a Kelvin wave should, therefore, be confined to the contour of the Grand<br />

Lac basin. Data from station G (see Figure 19) can be used to investigate this point.<br />

While the signal from the L1 mode is seen in the alongshore (east-west) component at<br />

A1, it appears as south-going currents at G (Figure 23). Thus the Kelvin wave has<br />

turned, following the deep basin contour of the Grand Lac instead of continuing to move<br />

along shore into the Petit Lac.<br />

Evidence for a Kelvin <strong>seiche</strong> response can also be seen in the spectra in Figure 21<br />

where the energy level of the along shore component is clearly dominant in the Kelvin<br />

<strong>seiche</strong> range.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

3.2.2.2 Mode T1<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 30 of 92<br />

For the only other strongly expressed mode, the first transverse mode T1 (n=10), the<br />

results of the cross spectral analysis are given in Figure 24, in that case for cross-basin<br />

stations 3 and 9. For the mode T1, the amphidromic pattern (Figure 23) displayed coherence<br />

between stations, which are part of the same or neighbouring amphidromic<br />

cell, but no coherence between stations in distant cells.<br />

Figure 24: The first transverse mode T1 during the summer interval 4 June to 24 Aug. 1950: top<br />

panel, structure and period predicted by the model. Bottom panels, coherence and phase between<br />

stations 3 and 9 from spectral analysis of SFE records of surface level fluctuations. Phase angle of<br />

+180° indicates a cross-basin oscillation. No coherence is found for station pairs which are not in<br />

the same amphidromic cell.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

3.2.3 Evidence of Poincaré waves<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 31 of 92<br />

Is this T1 mode a cross basin <strong>seiche</strong> or is it a Poincaré wave? The fact that a wave of<br />

near 11 h period has been detected in the shore-based water level records would favour<br />

interpretation as a standing cross-basin wave in the central part of the lake. However<br />

since a constant periodicity appears in the central part of the basin, independent of<br />

local topography, a Poincaré wave interpretation may be more likely. This is supported<br />

by results from the numerical model (Bäuerle, 1985), which predicts clockwise rotation<br />

in the amphidromic cells in that part of the lake.<br />

Poincaré waves are clearly seen in Figure 23 superimposed on the Kelwin wave pattern,<br />

particularly at the offshore station A3 and appear to be excited together with the<br />

Kelwin waves. To investigate this point further, the current records at station S in the<br />

central basin were examined in detail.<br />

3.2.4 Progressive vector analysis<br />

Progressive vector analysis was carried out for current components measured at station<br />

S in order to better understand the pattern described above. Figure 25 shows the vector<br />

diagram for the full record. This is a predominantly eastward oriented transport and the<br />

total running length is about twice that of the east-west length of the central bottom<br />

plateau. From the time markers it is obvious that there are periods of relatively slow<br />

transport as would be expected during summer stratification.<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

-1<br />

Lake Geneva, midlake station, depth 310m, from 6 June 14:00 to 12 October 24:00, 2001<br />

instrument depth 304 m<br />

marker every 24 h<br />

-3 0 3 6 9 12 15 18 21<br />

east west displacement (km)<br />

Figure 25 : Progressive vector diagram of currents measured at station S. For station location see<br />

Figure 19<br />

Frequently there are undulating sections in the progressive vector curve in Figure 25<br />

with two waves occurring between two 24 h markers. This obviously reflects a periodic<br />

oscillation of about 12 h clearly seen in the analysis above. The cusped motion pattern<br />

is evidently a superposition of a circular motion with a linear motion with a mean speed


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 32 of 92<br />

which is greater than the rotating vector. To learn more about the rotating motion, an<br />

excerpt of the progressive vector diagram has been produced with Figure 26. From this<br />

figure it is obvious that circular motions with periods close to 12 h are always executed<br />

in a clockwise sense of rotation.<br />

2.40<br />

2.15<br />

1.90<br />

1.65<br />

1.40<br />

Lake Geneva; midlake station, depth 310m, from 1 Sept. 14:00 to 5 Sept 14:00, 2001<br />

325<br />

343<br />

307<br />

379 397<br />

361<br />

415<br />

289<br />

253<br />

271<br />

433<br />

469<br />

451 487<br />

235<br />

505512<br />

12.75 13.00 13.25 13.50 13.75 14.00 14.25<br />

217<br />

199<br />

181<br />

163<br />

145<br />

91109<br />

127<br />

instrument depth 304m<br />

marker every 3 h<br />

73<br />

55<br />

1 37<br />

east west displacement (km)<br />

Figure 26 : Progressive vector plot for selected period of the trajectory shown in Figure 25.<br />

Further support for the Poincaré wave concept comes from spectra calculated by<br />

Mortimer et al. (1983) for current and temperature data at different stations in the central<br />

part of the lake. A statistically significant peak at the period of the T1 mode is always<br />

clearly seen in the temperature data at stations offshore (but rarely at the nearshore<br />

stations). However, it has to be realized that the T1 mode has Poincaré wave<br />

characteristics only in the central part of the Grand Lac basin.<br />

3.2.5 Mixing by internal <strong>seiche</strong>s<br />

During the period of summer stratification, <strong>mixing</strong> in lakes can be caused by vertical<br />

<strong>mixing</strong>, boundary <strong>mixing</strong>, interbasin density currents and river inflows. In Lac Léman,<br />

river inflows and interbasin density currents are of little importance during summer<br />

stratification. Vertical <strong>mixing</strong> can be related to wind induced shear or to internal <strong>seiche</strong><br />

and internal wave activity.<br />

Turbulent <strong>mixing</strong> can be expressed in terms of turbulent <strong>mixing</strong> coefficients, Kz, in<br />

analogy to molecular diffusion coefficients. This coefficient can be determined by the<br />

flux gradient method from integral changes of a tracer, such as temperature, over a<br />

certain time and a certain depth in the following way (Jassby and Powell, 1975):<br />

Kz =−1/ ( [ A(z)∆T<br />

∆z]<br />

) A(z') ∆T<br />

∆t dz'<br />

zmax �z<br />

19


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 33 of 92<br />

A(z’) is the lake surface at depth z’, T is temperature, t is time, z and zmax is the bottom<br />

of the lake. The flux gradient method is valid during periods when lakes warm up but is<br />

not valid when convective cooling overwhelms wind-induced <strong>mixing</strong>. In the present<br />

<strong>study</strong> we will apply this equation to temperature profiles taken in the central part of the<br />

Grand Lac basin in order to investigate the origin of vertical <strong>mixing</strong> in Lac Léman. This<br />

selection will minimize the effect of thermocline tilt due to internal <strong>seiche</strong>s and wind setup.<br />

The calculation of the vertical <strong>mixing</strong> coefficient Kz were carried out based on a set of<br />

monthly temperature profiles taken between 1987 and 1991 for the warming period<br />

from May to September. For each month, average profiles were established from the<br />

data by interpolating to the same depth intervals and eliminating the linear trend due to<br />

progressive longterm warming of the lake. Details are given in Michalski and Lemmin<br />

(1995). Results of this calculation are shown in Figure 27.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 34 of 92<br />

Figure 27 : Logarithm of the vertical turbulent <strong>mixing</strong> coef. Kz vs. logarithm of the Brunt-Väisälä<br />

frequency N 2 . The data are plotted as depth profiles with the surface at right and the bottom at left.<br />

The corresponding depths are indicated on each curve. A straight line is drawn for the section of<br />

each profile where the relation Kz ∝a(N 2 ) b can be validated. (a) for May (b = -0.5). (b) for the<br />

whole warming season from May to September (b = –0.4). (c) for a multiannual trend (b = –0.6).<br />

The hypothesis that the <strong>mixing</strong> coefficients result essentially from vertical <strong>mixing</strong> has<br />

been examined by looking at the correlation between Kz and the Brunt Vaisala frequency<br />

N. The Brunt Vaisala frequency is defined as<br />

N 2 =−(dρ/ dz)(g /ρ)<br />

where g is the gravitational acceleration. As in most lakes, the vertical density gradient<br />

is controlled by the temperature gradient dT/dz. The conversion from temperature into<br />

density is carried out using standard formulae. Different theoretical models exist for the<br />

relation between Kz and N2 of which all have the form<br />

Kz ∝a(N 2 ) b


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 35 of 92<br />

where a and b are constants. Using dimensional analysis, Welander (1968) suggested<br />

that, depending on the origins of turbulence, two limiting cases given by b = -0.5 for<br />

shear induced turbulence and b = -1 for cascading 2D turbulence. Heinz et al. (1990)<br />

summarized results from different lakes where b was found to range between – 0.4 and<br />

–0.7 indicating <strong>mixing</strong> in lakes is predominantly generated by local shear and internal<br />

waves. In Lac Léman, values of b are found between –0.4 and –0.6 (Figure 27) indicating<br />

that energy cascading is not important. A value of –0.4 falls outside the range<br />

predicted by Welander. However, Jassby and Powell (1975) who found the same value<br />

already noted that the assumptions of steady state and horizontal homogeneity made<br />

by Welander are most likely not fulfilled over periods of several months.<br />

Our results indicate that during summer stratification, turbulent <strong>mixing</strong> in the upper water<br />

column is predominantly caused by processes related to internal <strong>seiche</strong>s and progressive<br />

internal waves (Lemmin et al., 1998; Thorpe et al. 1996; Thorpe and Jiang,<br />

1998) quantified by the order of magnitude of the <strong>mixing</strong> coefficient. The increase of Kz<br />

with N 2b (and depth) ends at a depth of ≈ 90 m (see Figure 27). Below that depth an<br />

exponential correlation between Kz and N 2b cannot be established indicating that processes<br />

other than those considered by Welander dominate the turbulent <strong>mixing</strong>.<br />

3.3 NUMERICAL SIMULATIONS<br />

3.3.1 Model Description<br />

We use a slightly modified version of the three-dimensional numerical code ‘GETM’,<br />

developed by Burchard and Bolding (2002). A detailed mathematical description of this<br />

model is given in their report. Here we only mention its main features.<br />

The model solves the three-dimensional shallow-water equations for momentum and<br />

heat with a free surface as the upper boundary condition. The Boussinesq assumption<br />

has been adopted, implying that the balance of mass simplifies to a statement of zero<br />

divergence of the velocity field. As in all models of this type, pressure is computed from<br />

the hydrostatic balance, and certain types of barotropic and baroclinic waves cannot be<br />

reproduced. Among them are waves short in comparison with the local water depth and<br />

non-linear solitary waves. However, in the context of this section, <strong>mixing</strong> caused by long<br />

internal waves is emphasized, and these restrictions are hardly relevant.<br />

Quite to the contrary, the capability of the model to predict the evolution and propagation<br />

of steep non-linear waves or ‘bores’ is relevant in Lac Léman as illustrated below.<br />

To capture the essential physics of these waves, it is necessary to retain the non-linear<br />

advection terms in the horizontal momentum balance. Since it is well-known that the<br />

classical first-order upstream schemes for the numerical discretisation of these terms<br />

have great difficulties in reproducing steep waves because of their excessive numerical


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 36 of 92<br />

diffusion, we discretise the advection terms with so-called Total Variation Diminishing<br />

(TVD) schemes, as described in Burchard and Bolding (2002).<br />

The size and stratification of all lakes considered in the EUROLAKES projects suggests<br />

that the effects of the rotation of the earth cannot be neglected, and thus the Coriolis<br />

force is included in the model.<br />

The model equations are discretised on a staggered Arakawa-C grid in the horizontal<br />

coordinate, using generalised sigma-coordinates in the vertical. Due to the extremely<br />

steep topography of Lac Léman and the use of sigma-coordinates, great care has to be<br />

taken in computing the internal pressure-gradient, the driving force for the internal<br />

waves discussed in this section. We alleviate these problems using high horizontal<br />

resolution (250 m) and a special discretisation technique for the internal pressuregradient<br />

(Burchard and Petersen, 1997). The vertical resolution is 40 sigma levels,<br />

which is about the minimum to resolve the turbulent structure in the boundary layers<br />

and around the thermocline.<br />

Evidently, because <strong>mixing</strong> by internal waves is to be investigated here, a good turbulence<br />

model is required. We use a so-called Algebraic Reynolds Stress model (Canuto<br />

et al., 2001), solved in connection with two differential transport equations for the turbulent<br />

kinetic energy, k, and the specific dissipation rate, ω (see Umlauf et al., 2002).<br />

3.3.2 Mixing by Long <strong>Internal</strong> Waves - Model Results<br />

We investigated the activity of long internal waves at two positions in the lake which are<br />

dynamically very different and serve as two extreme cases spanning the range of <strong>mixing</strong><br />

activity in Lac Léman. The two positions are marked by yellow dots in Figure 28.<br />

Position A is located near the bottom of the deep plateau of the main basin. Measurements<br />

at this point have been discussed above. Position B is located at the entrance of<br />

the ‘Petit Lac’, the elongated south western appendix of Lac Léman. Measurements<br />

and model results at these two locations will be considered below.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 37 of 92<br />

Figure 28: Smoothed topography of Lac Léman used as the input for the numerical model. The<br />

measuring positions are marked by yellow dots. The geographical orientation of the lake has been<br />

turned 17 degrees in the clockwise direction on this and the following plots.<br />

The computation of the possibly complicated structure of the wind field over Lac Léman<br />

with a meteorological model was not part of the EUROLAKES project. Therefore, we<br />

had to force our model with highly idealized winds. We used homogeneous wind fields<br />

from SW (along the basin) and from NE (across the basin). These fields can be thought<br />

of as a first-order approximation of the well-known winds ‘Vent’ and ‘Bise’, respectively,<br />

which typically last only for a few days. Evidently, the current pattern close to the surface<br />

will exhibit large errors due to these simplistic wind fields. However, the excitation<br />

of basin-scale internal waves and their decay due to <strong>mixing</strong>, the main topic of this section,<br />

should be reproduced with much better accuracy. All runs have been initiated with<br />

zero velocities and an idealised typical late summer stratification with a well mixed<br />

epilimnion of 18 °C, a homogeneous hypolimnion of 5 °C, and a transitional thermocline<br />

at about 20 m depth.<br />

3.3.2.1 Mixing at Deepest Part of Lac Léman<br />

The measurements at point A (marked as S in Figure 19) near the bottom of the deepest<br />

part of the main basin have been discussed above. The position of the spectral<br />

peaks and the clockwise rotation of the velocity vector have been interpreted as strong<br />

indicators for the dominance of Poincaré waves (in winter possibly degenerated to simple<br />

inertial oscillations) at this location. In summer, considerable energy was also found<br />

in the along-basin component of the velocity at much lower frequency. Current speeds<br />

were demonstrated to be weak at all times and of the order of 1 to 3 cm/s. Additional<br />

measurements showed that the temperature profile in the lowest 50 meters of the water<br />

column is weakly unstable, most likely due to the geothermal heat flux. This thermal instability,<br />

however, is always compensated by the much stronger stabilizing effect of salinity<br />

in this region. These observations point towards very weak or negligible <strong>mixing</strong><br />

near the bottom, even though a definite conclusion has to await direct measurements of<br />

turbulence.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 38 of 92<br />

To investigate the ability of the model to reproduce these results, we started our runs<br />

with a typical late summer stratification for Lac Léman. The response to both idealized<br />

wind fields mentioned above, each with a duration of 1 day and a wind speed roughly<br />

corresponding to 7 m/s, was analysed at point A.<br />

v ( m / s )<br />

0 . 0 3<br />

0 . 0 2 5<br />

0 . 0 2<br />

0 . 0 1 5<br />

0 . 0 1<br />

0 . 0 0 5<br />

- 0 . 0 0 5<br />

- 0 . 0 1<br />

- 0 . 0 1 5<br />

0<br />

u<br />

S i m u l a t e d b o t t o m s p e e d f o r " V e n t "<br />

v<br />

- 0 . 0 2<br />

0 1 2 3 4 5<br />

t ( d a y s )<br />

v ( m / s )<br />

- 3<br />

x 1 0<br />

S i m u l a t e d b o t t o m s p e e d f o r " B i s e "<br />

1 0<br />

- 2<br />

- 4<br />

- 6<br />

8<br />

6<br />

4<br />

2<br />

0<br />

v<br />

u<br />

- 8<br />

0 1 2 3 4 5<br />

t ( d a y s )<br />

Figure 29: Time series of the x- and y-components u, v of the velocity at point A near the bottom.<br />

Left panel: with along-basin wind ‘Vent’. Right panel: with cross-basin wind ‘Bise’. Note the different<br />

velocity scales.<br />

The velocity components in x- and y-direction (see Figure 28) near the bottom are displayed<br />

in Figure 29 for both wind fields. It is evident that the cross-basin v-component is<br />

dominated by oscillations of roughly 10 hours in both cases. In particular for the ‘Bise’<br />

situation (right panel), it is clearly visible that these oscillations are correlated to similar<br />

oscillations in the along basin u-component with a respective phase shift of 90 degrees,<br />

a clear indication for Poincaré wave activity. This correlation can also be seen, though<br />

less clearly, for the ‘Vent’ situation, where the u-component is dominated by a low frequency<br />

contribution with a ‘period’ of approximately 3 days. Spectral analysis (not<br />

shown) confirms these results, in particular the peak around 10 hours and the higher<br />

energy at low frequencies in the u-component for the ‘Vent’ forcing. Current speeds associated<br />

with the higher frequency motions are at most 1 cm/s, and only the low frequency<br />

contribution of the along basin current can reach 2-3 cm/s for the relatively<br />

strong winds applied here.<br />

We conclude that the numerical model is able to correctly predict the most important<br />

components of the currents at this location. The somewhat shorter period of the Poincaré-waves<br />

is due to the fact that the model stratification in this example was slightly<br />

stronger than the measured one, leading to higher wave speeds and shorter periods. In<br />

addition, the measured spectra (Figure 21) represent an average over many weeks with<br />

periods of varying stratification, and perfect agreement cannot be expected.<br />

The turbulence model predicts zero turbulence at the bottom. Even though this is in apparent<br />

agreement with the absence of a measured well-mixed layer near the bottom<br />

(see above), this model result should not be overemphasized: It is well known that the<br />

class of turbulence models we adopted is not suited for the prediction of laminar-


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 39 of 92<br />

turbulent transition and strongly intermittent turbulence at very low Reynolds numbers.<br />

Both effects, however, have to be expected at the location investigated. At present,<br />

there exists no satisfactory theory for this regime of turbulent flows, and only when the<br />

boundary layer is fully turbulent, Reynolds stress models as that use here can yield<br />

reasonable results even when turbulence is rather weak. This has been shown recently<br />

by Lorke et al. (2002) for the bottom boundary layer of a small lake.<br />

3.3.2.2 Mixing at the Entrance of the Petit Lac<br />

The dynamics of mean currents and turbulence at point B, the entrance of the S-W appendix<br />

(‘Petit Lac’, see Figure 28), is quite different. Temperature profiles measured at<br />

this point for late summer and early winter 1987 are displayed in Figure 30. For the<br />

summer period (August/September), the temperature profiles exhibit a typical structure,<br />

which is also found in other years:<br />

• a well-mixed upper layer of about 10 m thickness<br />

• a strongly stratified upper thermocline from about 10 m to 30 m<br />

• a weakly stratified lower thermocline from about 30 m to 50 – 60 m<br />

• a well-mixed bottom boundary layer of 10 to 15 m thickness<br />

Later in the year, the thermocline is slowly mixed downward and eroded by penetrative<br />

convection, as is particularly visible from the slightly unstable temperature profile in December<br />

(see Figure 30). Finally, starting from January, the whole water column at this<br />

point is well mixed until re-stratification starts in spring (not shown). Current records at<br />

the entrance of the Petit Lac for a period of the same year 1987 are plotted in the right<br />

panel of Figure 30.<br />

d e p t h ( m )<br />

- 1 0<br />

- 2 0<br />

- 3 0<br />

- 4 0<br />

- 5 0<br />

- 6 0<br />

- 7 0<br />

0<br />

1 6 D e c<br />

T e m p e r a t u r e a t t h e e n t r a n c e o f t h e P e t i t L a c<br />

1 7 N o v<br />

2 8 O k t<br />

1 0 S e p<br />

1 8 A u g<br />

- 8 0<br />

5 1 0 1 5 2 0 2 5<br />

t ( d e g C )<br />

v ( m / s )<br />

0 . 4<br />

0 . 3<br />

0 . 2<br />

0 . 1<br />

- 0 . 1<br />

- 0 . 2<br />

- 0 . 3<br />

- 0 . 4<br />

0<br />

S p e e d a t t h e e n t r a n c e o f t h e P e t i t L a c<br />

6 0 m<br />

- 0 . 5<br />

0 5 1 0 1 5 2 0 2 5 3 0 3 5 4 0 4 5<br />

d a y s f r o m 2 1 . O k t . 8 7<br />

Figure 30 : Left panel: profiles of temperature at the entrance of the Petit Lac for the second half of<br />

1987. Right panel: measured in- and outflow velocities at the entrance of the Petit Lac at 10 m and<br />

60 m depth. Outflow is positive.<br />

1 0 m


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 40 of 92<br />

It is clearly visible, that the speeds at this point are at least one order of magnitude<br />

higher than at the bottom of the main basin. In addition, currents measured in the<br />

epilimnion and the hypolimnion close to the bottom are of comparable magnitude, but<br />

almost always of opposite sign causing a strong shear across the thermocline. It is to<br />

be noted, that waves of fluid entering and leaving the Petit Lac can be non-linear. An<br />

extreme example is the current reversal in the lower and upper layer at day 25 (see<br />

Figure 30): Currents in both layers change sign within less than 30 minutes (the resolution<br />

of our measurements), indicating a strongly non-linear internal ‘bore’ entering the<br />

Petit Lac. It is very likely that turbulence caused by these processes is crucial for shaping<br />

the vertical profiles of passive and active scalars in the water column.<br />

We tried to model the basic features at position B by forcing our model with a strong<br />

‘Vent’-type along-basin wind of about 7 m/s, which lasts for one day and was then<br />

switched off. Numerical studies showed that a ‘Bise’ wind event also causes high velocities<br />

at the entrance of the Petit Lac. However, turbulence characteristics and vertical<br />

shear are quite similar in both cases, and we considered it sufficient to look only at one.<br />

The left panel of Figure 31 illustrates the structure of the velocity field 38 hours after the<br />

wind has started (and terminated after 24 hours). This is precisely the time at which the<br />

return current, driven by the pressure gradient caused by the interface set-up, reaches<br />

the entrance of the Petit Lac.<br />

As can be seen from this figure, the return current occurs in the form of a strongly nonlinear,<br />

coastally trapped ‘bore’ in the Northern part of the entrance, and in form of a<br />

more gradual adjustment in the Southern part. After the shock wave has passed, the<br />

inflow currents of warm surface water into the Petit Lac are distributed fairly homogeneous<br />

across the entrance (not shown). As discussed above, the existence of such<br />

non-linear waves is also indicated by our current measurements, if the initial wind forcing<br />

is strong. The correct prediction of these waves depends to a large degree on the<br />

ability of the model to reproduce strong horizontal gradients in scalar and vector fields,<br />

i.e. on the quality of the advection scheme.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 41 of 92<br />

Figure 31: Left panel: surface velocities at the entrance of the Petit Lac (cf. Figure 28 for the geometry)<br />

38 hours after the start of an along-basin wind lasting for 24 hours. Dark red arrows indicate<br />

‘warm’ water, light red/orange arrows ‘cold’ water. Right panel: vertical velocity profile at<br />

point B (entrance of the Petit Lac) 50 hours after the start of the wind.<br />

The modelled vertical velocity profile 50 hours after the start of the wind, long after the<br />

‘bore’ of the return current has passed, is illustrated in the right panel of Figure 32.<br />

Since no measured velocity profiles are available at point B, we remark only on the<br />

most evident features of this profile, namely a strong shear near the bottom, a sharp<br />

maximum at about 47 m depth, a strong shear across the thermocline, and a more or<br />

less well-mixed upper layer.<br />

Figure 32: Left panel: vertical profile of the modelled temperature at point B (entrance of the Petit<br />

Lac) 50 hours after the start of the wind. Right panel: same as left panel, but now the turbulent diffusivity<br />

of momentum is displayed.<br />

The temperature profile at the same time is shown on the left panel of Figure 30. Remarkably,<br />

the vertical structure of this profile is very similar to that measured at the<br />

same location in late summer (see above): One easily identifies a well-mixed layer near


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 42 of 92<br />

the surface and the bottom, a strongly stratified upper thermocline, and a weakly stratified<br />

lower thermocline. Also the vertical extent and position of these different zones<br />

closely correspond to the measurements.<br />

The physical processes that shaped the thermocline structure become evident from the<br />

right panel of the same figure, which displays the turbulent diffusivity computed by the<br />

Reynolds stress model at the same time. As can be seen from this figure, the wellmixed<br />

bottom boundary layer is caused by strong turbulent diffusivities in the lowest 15<br />

meters, driven by the strong velocity shear near in this region. At the region of the velocity<br />

maximum (at about 47 m, see above), this shear is zero and diffusivities drop to<br />

very small values. This happens exactly at the lower edge of the stratified region.<br />

Above, the shear becomes strong again, turbulent diffusivities increase and cause an<br />

erosion of the thermocline from below. The result can be found in both, the measured<br />

and the modelled temperature profiles. In this region, the gradient Richardson number<br />

predicted by the turbulence model is approximately 0.25 (not shown), and thus <strong>mixing</strong><br />

occurs at high efficiency. From 30 to 10 meters depth, stratification becomes too strong,<br />

and turbulence is completely suppressed by local buoyancy effects. Only in the upper<br />

10 meters, wind <strong>mixing</strong> is strong enough to create a well-mixed region. Note, that at 50<br />

hours as in Figure 30 the wind has already been switched off, and diffusivities are only<br />

weak. This figure also gives a nice impression of the vertical numerical resolution required<br />

to resolve the basic features of the profile.<br />

3.4 CONCLUSIONS<br />

An analysis of internal <strong>seiche</strong>s dynamics was carried for Lac Léman combining field<br />

measurements and numerical modeling. Using field data of temperature, currents and<br />

surface elevation, it has been shown that only two modes of internals <strong>seiche</strong>s are sufficiently<br />

excited in Lac Léman to be considered significant. The first one is a Kelvin wave<br />

and the second one is a Poincaré wave. Model calculations have indicated that other<br />

<strong>seiche</strong> modes can only be excited by winds from certain directions. However, due to the<br />

topographic constraints particularly in the eastern part of the lake basin the wind field<br />

over the lake is strongly canalized and these winds do not exist in nature.<br />

The Kelvin wave progresses around the perimeter of the lake basin and its effects are<br />

most prominent in the nearshore zone. During the passage of the wave the thermocline<br />

descends by several meters and this provokes a thermocline displacement over the<br />

weakly sloping lateral zone which may easily reach 100 m. The combined alongshoredownslope<br />

velocity vector can reach speeds high enough to cause sediment erosion<br />

with potentially negative effects for the drinking water intake structures which are placed<br />

in the same zone. Since Kelvin waves are rather frequent in the Lake of Geneva, it can<br />

be expected that the dynamics of the nearshore zone vary on a periodic level during<br />

stratification.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 43 of 92<br />

As shown above, Poincaré waves are most prominent in the offshore regions of the<br />

central basin. However, water surface recorder records near the shore also indicated<br />

their presence. Their mayor axis is oriented in the cross-lake direction. The transport<br />

resulting from these waves will link the water masses in the center of the lake with the<br />

near shore zones. This will affect the water mass residence times, in particular shortening<br />

those in the central part of the basin. We also observed that oxygen concentration<br />

in the near bottom layers in the center of the lake fluctuates with the period of<br />

Poincaré waves. This is further indication of the link between the water masses in the<br />

center of the lake and the lateral zones. Due to their period of about 12 h during summer,<br />

a certain correlation in the forcing with the diurnal windfield over the lake (Lemmin<br />

and D’Adamo, 1996) can be envisioned.<br />

From our analysis of field studies of the longterm mean conditions of <strong>mixing</strong> it is indicated<br />

that internal <strong>seiche</strong>s are important in providing vertical <strong>mixing</strong>. Recently, it has<br />

been shown though that most of this <strong>mixing</strong> is actually generated in the near shore<br />

zone and then propagates into the open waters (Wuest et al., 2000). Thus, the interaction<br />

between near shore zones and the open water is also important for <strong>mixing</strong>. Furthermore,<br />

we have pointed to the importance of the interaction with the sloping sides of<br />

the lake and short progressive internal waves (Thorpe and Lemmin, 1999a, Lemmin et<br />

al., 1998). These waves and their breaking play a role in the production and redistribution<br />

of currents and stratification as well as <strong>mixing</strong> (Thorpe and Jiang, 1998). From our<br />

studies it appears that short progressive internal waves are often produced in the passage<br />

of non-linear internal <strong>seiche</strong>s (Thorpe et al., 1996).<br />

A state-of-the-art numerical model for the three-dimensional shallow-water equations<br />

has been compared to measured currents at two dynamically very different locations in<br />

Lac Léman: A low-energy point close to the bottom in the deepest part of the lake, and<br />

a very active region at the entrance of the appendix ‘Petit-Lac’, both for a typical latesummer<br />

stratification. Even though the wind field was highly idealized, the basic components<br />

of the currents at both locations could be reproduced: The nearly linear Poincaré-wave<br />

signal with very low current speeds at the deepest point of the lake, and the<br />

pattern of fast inflow and outflow currents including effects of the non-linear return-wave<br />

at the entrance of the Petit Lac after the wind had stopped.<br />

Clearly, the basic theory of linear shallow water waves in stratified basins is known<br />

since many decades, and the existence of these waves in Lac Léman does not come<br />

as a great surprise. However, the ability of our model to predict these waves, in particular<br />

their non-linear steepening at the entrance of the Petit Lac, can be taken as an<br />

indication for the reliability of the numerical scheme.<br />

Much more important as the mere prediction of internal waves is the question of their<br />

contribution to the overall <strong>mixing</strong> in Lac Léman. This was the major topic of the work<br />

package, and new insight into the physics of this process has been obtained. It must,<br />

however, be noted that due to the low current speeds in the deepest part of the lake, no<br />

useful information about the turbulent characteristics in the deep hypolimnion can be<br />

expected from the turbulence model. Since our turbulence model is state-of-the-art in


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 44 of 92<br />

geophysical modelling, this deficiency points into the direction of more fundamental investigations<br />

on this topic. As long as no measurements of microstructure in the water<br />

column are available, it can not even be definitely concluded at what level (if at all) turbulence<br />

exists in the lowest quarter of the hypolimnion.<br />

Quite encouraging results have, however, been obtained from the comparison of<br />

measured and computed quantities at the entrance of the Petit Lac. The measured<br />

characteristics of the velocity time series (namely the non-linear wave) in the case of a<br />

strong wind event could be reproduced. In addition, the good agreement of the measured<br />

and computed structure of the temperature profile suggest that the turbulence<br />

model yields reasonable turbulent diffusivities. All these results indicate that the Petit<br />

Lac could serve as the main ‘mixer’ of hypolimnetic water in the whole basin, even<br />

though enhanced turbulence at the lateral boundaries of the lake may also play a role.<br />

The precise mechanisms of how <strong>mixing</strong> in the Petit Lac affects the main basin are not<br />

yet known. The following possibilities appear to be, however, the most reasonable:<br />

• mixed water from the Petit Lac enters the hypolimnion of the main basin by advection<br />

through the strong mean currents in this region<br />

• mixed water intrudes horizontally into the main basin, driven by the density differences<br />

due to differential <strong>mixing</strong><br />

• heavy water is ‘strained’ over lighter water and causes additional local <strong>mixing</strong><br />

due to static instability of the water column<br />

All processes could be identified in the model results. Experimental confirmation, however,<br />

would be required to confirm if these processes in fact occur in the lake, and to<br />

what extent each of them contributes to the overall <strong>mixing</strong> in Lac Léman.<br />

3.5 REFERENCES<br />

Bäuerle, E. (1985) <strong>Internal</strong> free oscillation in the Lake of Geneva. Ann. Geophysicae.<br />

2/3: 199-206.<br />

Bohle-Carbonell, M. (1986) Currents in Lake Leman. Limnol. Oceanogr. 31: 1255-1266.<br />

Bohle-Carbonell, M., and D. v. Senden (1990) On internal <strong>seiche</strong>s and noisy current<br />

fields- theoretical concepts versus observations. Large Lakes Ed. M. T. a. C. Seruya.<br />

Springer. 81-105.<br />

Burchard, H., and K. Bolding(2002) GETM – A General Estuarine Transport Model,<br />

EUR 20253 EN, European Commission Joint Research Center, 21020 Ispra, Italy<br />

Burchard, H., and O. Petersen (1997) Hybridisation between sigma and z coordinates<br />

for improving the internal pressure gradient calculation in marine models with steep<br />

bottom slope, International Journal for Numerical Methods in Fluids, 25, 1003-1023,<br />

Caloi, P., M. Migani, and G. Pannocchia (1961) Ancora sulle onde interne del lago di<br />

Bracciano e sui fenomeni ad esse collegati. Ann. Geofisica, Roma 14: 345-355.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 45 of 92<br />

Canuto, V. M., A. Howard, Y. Cheng, and M. S. Dubovikov ( 2001) Ocean turbulence I:<br />

One-point closure model. Momentum and heat diffusivities. Journal of Physical Oceanography,<br />

31, 1413-1426,<br />

Heinz, G., et al (1990) Vertical <strong>mixing</strong> in Ueberlinger See, western part of Lake Constance.<br />

Aquat. Sci. 52:256-268.<br />

Jassby A. and T. Powell (1975) Vertical patterns of eddy diffusion during stratification in<br />

Castle Lake, California. Limnol. Oceanogr. 20: 530-543.<br />

Kanari, S. (1984) <strong>Internal</strong> waves and <strong>seiche</strong>s, Lake Biwa. S. Hone, ed. , Junk. Dordrecht,<br />

pp. 185- 235.<br />

Kanari, S. (1975) The long-period internal waves in Lake Biwa. Limnol. Oceanogr. 20:<br />

544-553.<br />

Lemmin, U. (1987) The structure and dynamics of internal waves in Baldeggersee.<br />

Limnol. Oceanogr. 32: 43-61.<br />

Lemmin, U., and N. D'Adamo (1996) Summertime winds and direct cyclonic circulation:<br />

observations from Lake Leman. Ann. Geophysicae 14: 1207-1220.<br />

Lemmin, U., et al (1998) Finescale dynamics of stratified waters near a sloping boundary<br />

of a lake. Physical processes in lakes and oceans Ed. J. Imberger. Amer. Geophys.<br />

Un., Washington, DC, Coastal and Estuarine Studies, 54:461-474.<br />

Lorke, A., L. Umlauf, T. Jonas, and A. Wüst, Dynamics of turbulence in low-speed oscillating<br />

bottom-boundary layers of stratified basins, Environmental Fluid Mechanics,<br />

accepted 2002<br />

Michalski, J. and U. Lemmin (1995) Dynamics of vertical <strong>mixing</strong> in the hypolimnion of a<br />

deep lake: Lake Leman. Limnol. Oceanogr. 40: 809-816.<br />

Mortimer, C.H. (1955) Some effects of earth rotation on water movements in stratified<br />

lakes. Verh. Int. Ver. Limnol. 12: 66-77.<br />

Mortimer, C..H. (1963) Frontiers in physical limnology with particular reference to long<br />

waves in rotating basins. Proc. 5th Conf. Great Lakes Res. Div., Univ. Michigan, 9-42.<br />

Mortimer, C.H. (1993) Long internal waves in lakes: review of a century of research.,<br />

Special report, No. 42, Univ. Wisconsin-Milwaukee, Center for Great Lakes Studies,<br />

117 pp.<br />

Mortimer, C.H., et al (1984) <strong>Internal</strong> oscillatory responses of Lake Geneva to wind impulses<br />

during 1977/78 compared with waves in rotating channel models. Commun. Lab.<br />

Hydraul., Ecole Polytech. Fed. Lausanne, No. 50, 89 pp.<br />

Richter, E. (1897) Seenstudien. Pencks Geogr. Abh., Wien 6: 121-191.<br />

Saggio, A., and J. Imberger (1998) <strong>Internal</strong> wave weather in a stratified lake. Limnol.<br />

Oceanogr. 43: 1780-1795.<br />

Service fédéral des eaux; SFE (1954). Les dénivellations du lac Léman. Département<br />

fédéral des postes et des chemins de fer. Report, maps, figures.<br />

Sirkes, Z. (1987) Surface manifestations of internal oscillations in a highly saline lake<br />

(the Dead Sea). Limnol. & Oceanogr. 32: 76-82.<br />

Thorpe, S. A. and U. Lemmin (1999a). <strong>Internal</strong> waves and temperature fronts on<br />

slopes. Ann. Geophysicae 17: 1227-1234.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 46 of 92<br />

Thorpe, S. T., U. Lemmin, et al. (1999b). Observations of the thermal structure of a lake<br />

using a submarine. Limnol. Oceanogr. 44(6): 1575-1582.<br />

Thorpe, S. A. and R. Jiang (1998). “Estimating internal waves and diapycnical <strong>mixing</strong><br />

from conventional mooring data in a lake.” Limnol. Oceanogr. 43: 936-945.<br />

Thorpe, S. A., et al (1996) High frequency internal waves in Lake Leman. Phil. Trans.<br />

Roy. Soc. London A 354: 237-257.<br />

Thoulet, M. J. (1894) Contribution à l'étude des lacs des Vosges. Bull. Soc. Geographie<br />

15: 557-604.<br />

Umlauf, L., H. Burchard, and K. Hutter, Extending the k-omega turbulence model towards<br />

oceanic applications, Ocean Modelling, accepted 2002.<br />

Wuest, A., G. Piepke, et al. (2000). “Turbulent kinetic energy balance as a tool for estimating<br />

vertical eddy diffusivity in wind forced stratified waters.” Limnol. Oceanogr. 45:<br />

1388-1400.<br />

4 MAPS AND TABLES OF FREE INTERNAL SEICHES IN UPPER LAKE<br />

CONSTANCE FOR PRACTICAL USE<br />

4.1 OBJECTIVE<br />

The assessment of the local intensity of internal <strong>seiche</strong>s due to their variable structure<br />

is of considerable concern in applications of water management, certain water constructions<br />

and other executive measures, for which essential information on transient<br />

currents and corresponding water displacements is required. For instance, the spill of<br />

harmful substances, drift of lost bodies, <strong>mixing</strong> and dispersion in various limnological<br />

and hydrological problems may depend during the period of stratification occasionally<br />

strongly on internal <strong>seiche</strong>s. Their local effect in such cases can be estimated on the<br />

basis of adequately resolved graphical representations of their variable spatial intensity.<br />

As rough information it is often sufficient to get an idea on the local variation of potential<br />

activity by superposed internal <strong>seiche</strong>s of different order.<br />

To enable the expert community without resort to detailed knowledge of the physics of<br />

internal waves, the oscillations have to be displayed in a form easy to grasp by inspection.<br />

This was done in a different delineation than the structures usually shown in terms<br />

of wave parameters such as lines of equal range and phase, which give a condensed<br />

overview. Instead, the resolution into momentary wave stages and local amplitude<br />

variations with time has been presented providing better imagination and at the same<br />

time more refined quantitative information on the structure. The corresponding horizontal<br />

motion is given in field representations of ellipses of the rotating current vectors<br />

with indication of the sense of rotation and the zero phase position for a defined moment<br />

of the vertical elevation. Such a description consists of several diagrams for one<br />

mode and sums up to a collection of numerous illustrations as the number of modes<br />

represented increases. Despite this inconvenience the use is nevertheless facilitated,<br />

as the corresponding diagrams of each mode have the same scale and the thematic<br />

content has been drawn in the same graphically proper forms. This description has<br />

been compiled for the first 15 modes of Upper Lake Constance. A few examples are<br />

selected in the following to give an idea for practising.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 47 of 92<br />

The work had been completed by Bäuerle and Ollinger (1991) in collaboration with<br />

Hollan in a project of the ISF. Later refined calculations by Bäuerle resulted in more<br />

details but confirmed the main structures of the oscillations already calculated in this<br />

early approach. Thus the compilation from 1991 is still worthwhile in the present context<br />

and may also serve as an example for the other large lakes under consideration. In<br />

particular reference to Lake Leman, the calculation with the same model has been carried<br />

out for the first 12 modes also by Bäuerle (1985). The extension to a similar presentation<br />

was included for merely two modes, thus not allowing for scanning the local<br />

potential contribution of all the 12 modes. The usefulness of the chosen graphic description<br />

of internal <strong>seiche</strong>s has been demonstrated very early by Bäuerle and Hollan<br />

(1983) for the case of the fundamental and a transverse mode of Lake Tanganyika enclosed<br />

in the monography on the lakes of the warm belt by Serruya and Pollingher<br />

(1983). This reference is quoted within the preceding one. C. Serruya recommended in<br />

this context to carry out such work on other large stratified lakes as accomplished later<br />

in the advanced version for Lake Constance here.<br />

4.2 OUTLINE OF THE CALCULATION<br />

The strength of internal <strong>seiche</strong>s is most adequately described in terms of forced oscillations.<br />

However, this approach implies precise knowledge of the driving agent, which is<br />

mostly the wind field over the lake. The horizontal variation of this quantity is generally<br />

not sufficiently known for that purpose, in particular, for large lakes. Moreover, the superposed<br />

different internal modes excited during one event have to be decomposed for<br />

identification of the associated single contributions. Desisting from such a difficult description,<br />

which is practically beyond reach, the relative structures and natural periods<br />

in terms of free oscillations may already serve for essential information to quantify the<br />

phenomenon. If certain observations exist on the mean amplitude of internal <strong>seiche</strong>s<br />

with respect to typical wind fields and the stratification in the lake, the determination<br />

relative to an arbitrary factor may be converted to absolute values, which often suffices<br />

for assessment.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 48 of 92<br />

Figure 33 : Typical temperature stratification in the western part of Lake Constance (Lake Überlingen)<br />

mid October 1972 and the corresponding density stratification (relation ρ = ρ(<br />

T ( z))<br />

see<br />

text, from Bäuerle (1981)<br />

The problem is most conveniently solved by an eigenvalue-method, as the structures<br />

and periods of the oscillations are calculated as distinct constituents of each solution<br />

and result therefore very precisely. The question is, how detailed the hydrodynamic<br />

model is formulated to simulate nature. In the present context a two-layer model has<br />

been adopted including the Coriolis force. By the same reason, as the forcing has been<br />

kept out of concern, friction is not considered, since it represents generally a lakespecific<br />

process. It may be introduced by empirical information and rough linear assumption,<br />

if required for estimation. The approach under these conditions yields still a<br />

very useful description, when the calculations cover also characteristical stratifications<br />

during the warm season, what has been carried out for Upper Lake Constance. The<br />

simplification by a two-layer model restricts the solutions to the fundamental vertical order.<br />

This limitation is to a certain extent serious, as internal <strong>seiche</strong>s of second vertical<br />

order exist in lakes and should be included into the consideration. Since they are not so<br />

frequent and appear less pronounced generally than the fundamental vertical modes,<br />

their omittance may be tolerable for the time being.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

momentary surface<br />

h 1<br />

h 2<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

mean position<br />

of surface<br />

mean depth position<br />

of interface momentary depth<br />

position of interface<br />

density ρ 2<br />

density ρ 1<br />

Page 49 of 92<br />

Figure 34 : General sketch and definition of quantities of a two-layer model with respect to the<br />

stratification in a deep temperate lake in summer<br />

Lake Constance is a large deep lake in the moderate zone and develops therefore<br />

during the warm season a stratification, which consists of a relatively shallow surface<br />

layer (epilimnion) over a deep lower layer (hypolimnion). A typical example of the vertical<br />

density variation is given in Figure 33 for early autumn (October 1972). The relation<br />

between density and temperature applied here is valid for the waters of Lake Con-<br />

2<br />

stance (Hollan and Simons, 1978) and reads: ρ = ρ0<br />

−α<br />

( T − T0<br />

) with ρ0 = 1.000145<br />

g/cm³, T in °C, T0 = 4°C, α = 7.3 ·10 -6 g/(cm² °C²). While the upper depth range of 20 m<br />

is covered by the epilimnion, the hypolimnion extends from about 30 m to the maximum<br />

depth of 254 m, or on the average between 30 m and the mean depth of 100 m. The<br />

difference of thicknesses is even increased during summer, since the surface layer is<br />

generally less deep till October, when cooling becomes stronger. Such conditions are<br />

appropriate for a two-layer model of constant equivalent depth he . This quantity appears<br />

in the fundamental equations of internal <strong>seiche</strong>s in a two-layer system and allows for a<br />

description similar to surface (barotropic) <strong>seiche</strong>s, if it can be assumed as constant.<br />

With the definition of the two-layer model (see Figure 34) by an idealized step-like density<br />

stratification:<br />

�ρ1<br />

0 ≤ z < h1<br />

ρ0<br />

= �<br />

�ρ2<br />

h1<br />

≤ z < h(<br />

x,<br />

y)<br />

where x,y,z represent the coordinates of a Cartesian system with z directed vertically<br />

downward (z=0: surface), and ρ1, ρ2 constant densities with ρ2 > ρ1. The equivalent depth<br />

h e<br />

h1<br />

h<br />

=<br />

h + h<br />

1<br />

represents a quantity of low variation as to the deep and steep depth configuration of<br />

Lake Constance.<br />

It is therefore reasonable to approximate he by a constant value,


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

(1)<br />

h1<br />

h2<br />

=<br />

h + h<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

h e<br />

1<br />

2<br />

Page 50 of 92<br />

where h2 gives now a corresponding constant thickness of the lower layer. In contrast to<br />

the gravity g determining the hydrodynamic pressure forces of surface <strong>seiche</strong>s, the low<br />

density difference between both layers at the interface implies that the corresponding<br />

forces as to internal <strong>seiche</strong>s are governed by reduced gravity, i.e.:<br />

∗<br />

(2) g = g ⋅ε<br />

with<br />

ρ2<br />

− ρ1<br />

ε =<br />

ρ<br />

From this reason result relatively high internal amplitudes of several meters and considerably<br />

long periods of several hours up to several days compared to those of surface<br />

<strong>seiche</strong>s in the same basin. As example , the period T1 of the fundamental mode and the<br />

corresponding frequency ω1 in a rectangular lake of length L, constant equivalent depth<br />

he, without influence of the earth’s rotation, read according to Merian’s formula:<br />

(3)<br />

L<br />

T =<br />

∗<br />

g h<br />

2<br />

1<br />

e<br />

and<br />

= π<br />

ω1 2<br />

∗<br />

g he<br />

In (3) the term g he<br />

∗<br />

represents the phase velocity ci of long internal waves in this system,<br />

which is considerably lower than that of long surface waves cs = g H , with H =<br />

h1+h2. Since the periods of surface <strong>seiche</strong>s obey the corresponding relation as in (3)<br />

with respect to cs , the great difference is evident.<br />

The third physical parameter in the fundamental equations accounts for the influence of<br />

the earth’s rotation. The effective component of the rotational vector of the earth, the<br />

Coriolis parameter f, reads:<br />

(4) f = 2Ωsinϕ<br />

with Ω = 7.29 ·10 -5 s -1 the angular velocity of the earth and ϕ the geographical latitude.<br />

For the mean geographical latitude of Lake Constance, ϕ = 48°N, f amounts to<br />

1.07 ·10 -4 s -1 .<br />

The derivation of the governing equations and their numerical solution is treated by<br />

Bäuerle (1981). After separation of the sinusoidal time dependency, the system of<br />

equations is solved for distinct eigen-frequencies ωn (n = 1,2,3 ...). It describes the dynamical<br />

relations for the dependent variables of the lower layer, which are the amplitudes<br />

of the volume transport 2 V� (x,y) = (U2(x,y), V2(x,y)) and the amplitude of the vertical<br />

displacement of the interface ζ2(x,y) at the top of this layer. It reads:<br />

∗ ∂ζ<br />

2<br />

iω U 2 + fV2<br />

+ g he<br />

= 0<br />

∂x<br />

∗ ∂ζ<br />

2<br />

(5a) iω V2<br />

− fU 2 + g he<br />

= 0<br />

∂y<br />

∂V2<br />

iω ζ 2 + fU 2 + = 0<br />

∂y<br />

with the boundary condition:<br />

L


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

(5b) V ⋅ n = 0<br />

� �<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

2<br />

Page 51 of 92<br />

The condition (5b) means that there is no volume transport through to the rigid boundary<br />

expressed by the vanishing scalar product of 2 V� with the unit vector n � normal to the<br />

boundary. (5a) together with the condition (5b) forms an eigenvalue problem with an infinite<br />

number of distinct eigen-frequencies and corresponding eigen-solutions. The<br />

imaginary unit i = −1<br />

in the equations (5a) indicates that the solutions are given in<br />

�<br />

complex notation. Thus the eigen-solutions of the volume transports V 2,<br />

n (x,y) and the<br />

amplitudes ζ 2,<br />

n (x,y) represent complex functions and the corresponding real functions<br />

are evaluated for presentation in the course of the complex mathematical formulation.<br />

This treatment is also inherent to the nature of the oscillations as horizontally rotating<br />

long waves, which is caused by the influence of the Coriolis force. Due to this pattern a<br />

peculiar task of sufficiently resolved delineation for practical use has to be solved by a<br />

proper design, which is demonstrated in this report.<br />

A general property of the mathematical solution of (5a,b) is that the eigen-functions<br />

�<br />

V 2,<br />

n and ζ 2,<br />

n are determined except for a free factor. Thus the calculation yields the<br />

structure of the modes relative to 100% either of the maximum elevation of ζ 2,<br />

n or of the<br />

�<br />

maximum volume transport V 2,<br />

n encountered in the lake. Since the most interesting<br />

�<br />

functions for application with respect to the structure are ζ 2,<br />

n and V 1,<br />

n , which represents<br />

the amplitude of the volume transport in the upper layer, the relation of the latter quantity<br />

has to be supplemented here:<br />

� �<br />

(6) V 1,<br />

n = −V2,<br />

n<br />

From (6) the vertically averaged current in the epilimnion is inferred by<br />

�<br />

� V1,<br />

n<br />

(7a)<br />

v1,<br />

n =<br />

h<br />

while that in the lower layer results from<br />

�<br />

� V2,<br />

n<br />

(7b)<br />

v2,<br />

n =<br />

h<br />

with h2 given according to (1), if he and h1 had been prescribed.<br />

1<br />

2<br />

The equations (5a,b) are not solvable analytically even with constant he neither for very<br />

simple geometrical approximations of the basin nor for the irregular shape assumed<br />

here. Therefore, the solutions are determined by horizontal discretisation of the dependent<br />

variables and the rigid boundary. Their derivations in the fundamental equations<br />

(5a) are replaced by central differences of the variables discretised in a square<br />

grid of 1.4 km mesh size, as used by Hollan et al. (1980) for calculation of the surface<br />

<strong>seiche</strong>s of Lake Constance. The grid is shown Figure 35.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 52 of 92<br />

Figure 35 : Outline of the numerical grid of 1.4 km mesh size adopted for Upper Lake Constance<br />

(Rao - grid in Hollan et al. (1980))<br />

The variables U2,n, V2,n and ζ2,n are defined on this grid, each staggered by 700 m. The<br />

details of the numerical solution are treated by Bäuerle (1981). The minimum depth at<br />

the numerical grid points is assumed to be somewhat larger than the thickness h1 of the<br />

epilimnion . This problem is discussed in Bäuerle’s (1981) treatise in more detail.<br />

4.3 AUXILIARY FORMULATIONS FOR EVALUATION OF EIGEN-PERIODS WITH<br />

RESPECT TO DIFFERENT STRATIFICATIONS<br />

The advantage of the two-layer model with constant equivalent depth consists in the<br />

property that the eigen-periods determined for selected different stratifications may be<br />

transformed to the cases of other stratifications by simple auxiliary relations. In order to<br />

arrive at a concise representation of the eigen-periods, as they vary with the modal order,<br />

with stratification and with respect to the Coriolis effect, this convenience of the<br />

model is utilized as follows. The transformation of the eigen-periods to different stratifications,<br />

if they are computed for one case, is non-ambiguously given, when there is no<br />

influence of the earth’s rotation, f = 0. Such internal <strong>seiche</strong>s represent standing oscillations.<br />

Their period spectrum is determined exclusively by the configuration of the basin<br />

and differs for each peculiar stratification by a constant factor. This factor q is inferred<br />

from the phase velocities ci of long internal waves<br />

∗<br />

(8) ci<br />

= g he<br />

which is characteristic for each case of stratification, as mentioned in the context of<br />

equation (3).<br />

When the eigen-frequencies have been calculated for a selected stratification which is<br />

considered formally as reference stratification and are designated by ωn(ref), the conversion<br />

factor q for the corresponding frequencies ωn(novel) of another stratification is<br />

defined by the ratio:<br />

ci<br />

( novel)<br />

(9)<br />

q =<br />

c ( ref )<br />

and the relation of the eigen-frequencies reads:<br />

(10) ω ( ) = q ⋅ω<br />

( ref ) n = 1, 2, 3...<br />

n<br />

i<br />

novel n<br />

Since the graphical representation of the dependency on the stratification and the influence<br />

of the earth’s rotation is very lucid and simple at the same time, if the frequencies<br />

are normalised to the fundamental frequency ω1 for f = 0, the conversion relation (10)<br />

may be put alternatively into the form


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

(11) ω ( ) = p ⋅ω1(<br />

novel)<br />

n = 1, 2 ,3...<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

n<br />

novel n<br />

Page 53 of 92<br />

where the factors pn represent the normalised eigen-frequencies determined for one<br />

reference case of stratification by the relation (11), i.e. by pn = ωn(ref) / ω1(ref) for f = 0.<br />

The corresponding relations to (10) and (11) in terms of eigen-periods Tn read:<br />

(10)* ( novel)<br />

= T ( ref ) / q<br />

Tn n<br />

(11)* T n ( novel)<br />

= T1<br />

( novel)<br />

/ pn<br />

with pn = T1<br />

( ref ) / Tn<br />

( ref ) , resp.<br />

In order to make use of (11), merely ω1(novel) of the new stratification has to be determined<br />

from equations (9) and (10) for n = 1. It is this notation which had been adopted<br />

due to its feasibility for evaluation and delineation.<br />

The definition and evaluation for a reference stratification with f = 0 has been compiled<br />

in Table 1.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

mode ωn in 10 -5 sec Tn in hours<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

σ n = ωn<br />

1 1.910 91.38 1.00<br />

2 3.349 52.11 1.75<br />

3 4.699 37.14 2.46<br />

4 6.255 27.90 3.27<br />

5 7.587 23.00 3.97<br />

6 8.504 20.52 4.45<br />

7 9.096 19.19 4.76<br />

8 10.538 16.56 5.52<br />

9 11.223 15.55 5.88<br />

10 11.849 14.73 6.20<br />

11 12.836 13.60 6.72<br />

12 13.087 13.34 6.85<br />

13 14.016 12.45 7.34<br />

14 14.762 11.82 7.73<br />

15 15.430 11.31 8.08<br />

ω1<br />

Page 54 of 92<br />

Table 1: Eigen-frequencies, eigen-periods and normalised eigen-frequencies of the first 15 modes<br />

for an exemplary reference stratification with f = 0, ε = 5.3 ·10 -4 , he = 21m, h1 = 30m and<br />

h2 = 70m. This case has been observed in the western part of Lake Constance in October 1972 and<br />

had been assumed for the explanation of transverse internal oscillations in this region (Hollan<br />

1974, Bäuerle 1981)<br />

4.4 THE EIGEN-PERIODS INCLUDING THE CORIOLIS EFFECT<br />

As to the eigen-frequencies, the generally strong and complicated influence of the<br />

Coriolis force can be condensed into a very instructive and handy diagram for evaluation<br />

with the preceding consideration in mind. The effect of the earth’s rotation increases<br />

the larger the lake is and the smaller the phase velocity of the internal wave is.<br />

The latter may be due to weakening of the stratification or lowering the values of the<br />

equivalent depth by diminishing the thickness of the upper layer. The dependency on<br />

rotation is at best expressed with respect to the rotation number<br />

( 0)<br />

(12)<br />

F = f /ω<br />

1<br />

( 0)<br />

where F is the ratio of the local Coriolis frequency f and the eigenfrequency ω 1 of the<br />

fundamental mode without earth rotation.


σσ<br />

Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

σ<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0 5 10 15 20<br />

Figure 36 : Dependency of the dimensionless eigen-frequencies<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

F<br />

σn ωn<br />

ω<br />

Page 55 of 92<br />

= on the rotation<br />

( 0)<br />

1<br />

( 0)<br />

F = f /ω for 15 modes of internal <strong>seiche</strong>s in Upper Lake Constance on the basis of a<br />

number<br />

1<br />

two-layer constant equivalent depth model (Figure 35) The vertical dashed lines indicate the dates<br />

for which the calculations of the 1991 report were done. The circles refer to Figures 38, 41, 44, 47<br />

and Figure 50 through Figure 55.<br />

( 0)<br />

Since the stratification enters into ω 1 proportionally to the phase velocity according to<br />

(8), the rotation number F increases, when either the stratification as given in (2) or the<br />

equivalent depth he diminish. The latter is effectuated by reduction of h1, the thickness of<br />

the surface layer. As mentioned above, the influence of the size of the lake is, by the<br />

way, detected with the help of (3) in terms of the length L in a rectangular basin. In this<br />

case the rotation number is proportional to L.<br />

With constant Coriolis frequency the rotation number F depends (inversely) on the<br />

( 0)<br />

( 0)<br />

value of ω 1 . As discussed above, ω 1 decreases with the length of the basin and with<br />

decreasing phase velocity. In other words, the larger and the less stratified the lake is,<br />

the larger is the influence of the earth rotation.<br />

At low F the internal <strong>seiche</strong>s resemble standing oscillations. With increasing F the effect<br />

of rotation becomes dominant and the structure of the oscillations takes the form of a<br />

rotating wave propagating around amphidromic points, where the amplitude of vertical<br />

displacements is zero. The number of amphidromic systems for the lower order modes<br />

is generally identical with the modal number.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 56 of 92<br />

The influence of the rotation number on the eigen-frequencies is presented in condensed<br />

form in the diagram shown in Figure 36. Some explanations are necessary to<br />

make use of the information compiled in this figure. Increasing influence of rotation is<br />

represented by increasing rotation number F on the abscissa. On the other side, the or-<br />

( 0)<br />

dinate gives the normalised eigen-frequencies σ n = ωn<br />

ω1<br />

, discerned by increasing<br />

modal order up to 15. Thus, a family of 15 curves is displayed, each curve starting on<br />

the ordinate at F = 0, with the fixed normalised distribution for the case of Upper Lake<br />

Constance. These numbers σ n for f = 0 are listed in the fourth column of Table 1, as<br />

their general values result from the adopted reference stratification. For f = 0, they are<br />

the same for any different stratification as outlined in the previous section. These points<br />

on the ordinate mean the onset of the corresponding function showing the increasing<br />

influence of the earth’s rotation on the modal relative frequency in question.<br />

The effect of the stratification and varying equivalent depth is incorporated in the nor-<br />

( 0)<br />

malising fundamental frequency ω 1 . As to the other coordinate, the rotational number<br />

F gives the form, which reflects the dependency on the earth’s rotation with respect to<br />

the size and stratification of the lake. Since both variables are normalised to the same<br />

( 0)<br />

absolute quantity ω 1 , there is a simple method of evaluation with the help of this diagram.<br />

Given a particular stratification and equivalent depth, the corresponding rotation number,<br />

say Fe, is determined according to (12) and (3). The vertical line at this value on the<br />

abscissa crosses the curve family at 15 respective ordinate values σ n(<br />

F1<br />

) ,<br />

n = 1, 2, .., 15. From these relative eigen-frequencies the absolute values ωn(Fe) are<br />

( 0)<br />

obtained by multiplication with ω 1 . The corresponding eigen-periods result from the<br />

formula Tn ( F1<br />

) = 2π<br />

/ ωn<br />

( Fe<br />

) . This evaluation is presented for two examples selected from<br />

33 cases of stratification taken from observations in the lake during the period from<br />

1985 through 1989. The 15 eigen-periods for each case are tabulated by Bäuerle &<br />

Ollinger (1991) in their original report. The data of both examples shown here concern a<br />

case in spring on 13 April 1989 and another one in late summer on 30 August 1989.<br />

They represent a situation, which is strongly influenced by rotation (spring) and predominantly<br />

by stratification (late summer), respectively. The extracted stratification pa-<br />

( 0)<br />

( 0)<br />

rameters and resulting fundamental frequencies ω1 and periods T 1 , as well as the<br />

rotation numbers F are listed (Table 2).


13.4.<br />

1989<br />

1/q=<br />

3.3276<br />

30.8.<br />

1989<br />

1/q=<br />

0.7419<br />

Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

date h1 he<br />

13.04.1989 20 16.0<br />

30.08.1989 15 12.7<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

T1<br />

T2<br />

ρ1<br />

ρ2<br />

7.10 1.000075<br />

5.00 1.000138<br />

18.80 0.998546<br />

5.30 1.000133<br />

ε ci<br />

ω<br />

( 0)<br />

1<br />

( 0)<br />

1<br />

Page 57 of 92<br />

T F<br />

0.63 9.93 0.57 304.2 18.65<br />

15.86 44.54 2.57 67.8 4.16<br />

Table 2 : Approximation of temperature profiles from the central as well as deepest (254 m) position<br />

Fischbach-Uttwil of Upper Lake Constance in spring and late summer 1989 by two layers of<br />

constant density and resulting parameters of a two-layer equivalent depth model of internal<br />

<strong>seiche</strong>s. The total depth of the constant equivalent depth model is :<br />

H = h1+h2 = 100 m.<br />

date date of observation<br />

h1 depth of the upper layer [m]<br />

he equivalent depth [m]<br />

T1, T2 constant temperature of the upper and lower layer [°C]<br />

constant density of the upper and lower layer [g/cm³]<br />

ρ1, ρ2<br />

ε relative density difference ε = (ρ2 - ρ1)/ρ2 , [x10 -4 ]<br />

phase velocity of long internal waves [cm/s]<br />

ω eigen-frequency of the fundamental mode resulting from the two-<br />

ci<br />

( 0)<br />

1<br />

layer equivalent depth model for f = 0 [x10 -5 /s]<br />

( 0)<br />

T 1 corresponding eigen-period [h]<br />

( 0)<br />

F rotation number F = f / ω<br />

1<br />

F 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15<br />

18.65 443.2 206.7 140.6 104.8 83.6 70.7 60.6 53.7 48.3 43.4 39.5 37.1 34.7 32.7 30.5<br />

0 304.1 173.4 123.6 92.8 76.5 68.3 63.9 55.1 51.7 49.0 45.3 44.4 41.4 39.3 37.6<br />

4.16 72.9 41.7 28.5 21.3 17.5 15.1 12.9 11.2 10.7 10.5 9.6 9.3 8.7 8.3 8.0<br />

0 67.8 38.7 27.6 20.7 17.1 15.2 14.2 12.3 11.5 10.9 10.1 9.9 9.2 8.8 8.4<br />

Table 3 : Eigen-periods (in h) of the first 15 modes of internal <strong>seiche</strong>s in Lake Constance for stratification<br />

in spring (13 April 1989) and late summer (30 August 1989) displayed with and without<br />

(F = 0) the effect of the earth’s rotation.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 58 of 92<br />

Finally, the relative frequencies σ ( i ) with i = 1, 2 designating both cases, are read<br />

from the diagram in Figure 36. The absolute eigen-periods T ) resulting from<br />

n F<br />

T F ) = T / σ ( F ) are compiled in Table 3.<br />

n(<br />

i 1 n i<br />

In the spring case (13.4.1989) the periods T1 through T15, are rather long varying from<br />

443 h through 31 h, resp. Consequently the rotation number is relatively high,<br />

F1 = 18.65, and the influence of the Coriolis force is dominant. The effect is evident by<br />

the associated periods for the same case without rotation (f = 0) given in the second<br />

row for this date. The latter periods have been calculated from the internal phase velocity<br />

in this stratification quoted in Table 2, by using the relations (9) and (10)* with respect<br />

to the periods of the reference case listed in Table 1. There are considerable differences<br />

in the modal pairs of periods for f ≠ 0 and f = 0 in this case, which exhibit the<br />

strong rotational effects. It is remarkable that the eigen-periods of the lowest modes are<br />

considerably reduced for f = 0 compared to those influenced by rotation. For higher<br />

modal order from seven on the periods for f = 0 are higher. It is worthwhile to notice,<br />

that all the eigen-periods are greater than the inertial period. The consequences will be<br />

discussed in the following section.<br />

In the late summer case (30.8.1989) the rotation number (F2 = 4.16) is much smaller.<br />

While the periods vary from 72.9 h through 8.0 h for f ≠ 0 with the modal order from 1<br />

through 15, the corresponding periods for f = 0 deviate relatively less compared to the<br />

lowest orders of the previous case. The associated periods for f = 0 range from 67.8 to<br />

8.4. Despite that seemingly small effect on the eigen-periods, the effect of the earth’s<br />

rotation is of crucial importance for the higher modes. From general reasons, which are<br />

not delineated here, this is the case for the modes higher than order 5 which is indicated<br />

in Figure 36 by the straight line from the origin, σ = F, and the modes above it at<br />

F2 = 4.16. This transition is also reflected in the shift of relative amount of the periods<br />

between modal orders 5 and 6 given in Table 3. for f = 0 and f ≠ 0 in the last two lines,<br />

respectively.<br />

A final remark is in order for the application in case of an observed stratification which is<br />

not represented in the tabulated cases for f ≠ 0. Such a situation is generally to be expected<br />

and the large amount of 33 calculated examples has been achieved to meet this<br />

problem as to the periods. As approximation it is considered to be sufficient, if the most<br />

similar calculated case is found out as to the interesting stratification. If more precise<br />

assessment is required, an interpolation between two adjacent calculated cases may<br />

serve in the way that ci of one of them is greater and of the other is lower than that of<br />

the stratification in question.<br />

4.5 THE HORIZONTAL STRUCTURES OF THE INTERFACE AMPLITUDES<br />

Progressive rotating waves are usually illustrated in a diagram, which shows corange<br />

and so-called cotidal lines representing lines of equal amplitudes and of equal phases,<br />

resp. This delineation is inherent to the mathematical expression of such eigenoscillations.<br />

For a given eigen-frequency ωn the corresponding amplitude function of the<br />

solution as to the interface elevations reads:<br />

(13) ζ ( , y,<br />

t)<br />

= A ( x,<br />

y)<br />

cos( ω t + ϕ ( x,<br />

y))<br />

n = 1, 2, ...<br />

n x n<br />

n n<br />

n( Fi


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 59 of 92<br />

Since free eigen-oscillations are determined except for an arbitrary constant factor, the<br />

amplitudes An(x,y) are merely known to their relative distribution. Their absolute maximum<br />

may be ascribed the value of 100%, which occurs for every modal order generally<br />

at a different position in the interface. The corange lines show the horizontal structure<br />

and are displayed by equal percentage increments. The evaluation of the solution (13)<br />

is in terms of the formula<br />

(13a) A n ( x,<br />

y)<br />

= Pk<br />

with Pk in steps of 10% from -100% through +100%, correspondingly identified by the<br />

index k in the range {-10 (1) 10}.<br />

A<br />

B D<br />

C<br />

G<br />

F<br />

E<br />

Figure 37 : Contour of the numerical grid of Upper Lake Constance shown in Figure 35with the 19<br />

sites, from where the time dependent amplitude variation is illustrated in Figure 40, Figure 43,<br />

Figure 46 and Figure 49.<br />

Superposed in the same diagram are usually the cotidal lines, which represent the process<br />

of wave propagation. They are evaluated in equal steps of phase increase in the<br />

argument of the cosine function in (13). If the maximum amplitude is considered as<br />

phase stage, the cotidal lines are described by the formula<br />

2π<br />

k<br />

(13b) + ϕ n ( x,<br />

y)<br />

= 0 with k = 0, 1, 2, ..., K,<br />

K<br />

where K is the integer which divides the wave cycle into equal phase steps of 2π / K .<br />

The general pattern of cotidal lines in horizontally rotating waves is the radial arrangement<br />

of the curves meeting another in amphidromic points. When the wave progresses<br />

in the mathematical positive sense, i.e. is turning around the amphidromic point in anticlockwise<br />

direction, this point and system is called cyclonic, as the sense of the earth’s<br />

rotation is the same. The opposite sense of wave propagation designates the amphidromic<br />

system as anticyclonic.<br />

In the following, four sets of diagrams are presented as a selection from a total number<br />

of 30 which have been compiled for consultation of structural details of the first 15<br />

modes. The doubling results from the elaboration for the two aforementioned contrasting<br />

stratifications, namely, the case dominated by rotation in spring time (13 April 1989)<br />

and the case of governing gravity effects in late summer (30 August 1989). Particularly,<br />

the first and ninth modes are presented here, because the pattern of the structures and<br />

their variation with increasing modal order are disclosed well with this choice. Moreover,<br />

J<br />

I<br />

H<br />

L<br />

K<br />

M<br />

O<br />

N<br />

P<br />

R<br />

Q<br />

S


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 60 of 92<br />

they give a good idea of the instructive and feasible survey possible by the structurally<br />

more resolving illustrations.<br />

Before entering into the graphical inspection a reference map of 19 adjacent places on<br />

the boundary and in the interior of the lake has to be considered. These selected<br />

points, designated by the letters from A through S, are shown in Figure 37 on and inside<br />

the contour of the numerical grid. They serve for a proper display of local time histories<br />

of the interface displacements and are referred to in this peculiar thematic diagram<br />

in the sequence of illustrations for each selected mode. Besides these examples,<br />

there is another subset of diagrams, which shows transverse and longitudinal sections<br />

of the structure of interface displacements throughout the lake or in bays. Although<br />

these illustrations provide another comprehensive information, they are only mentioned<br />

here for the sake of brevity.<br />

After this preliminary note, we resort to the standard representations by corange and<br />

cotidal lines as outlined above. The corresponding diagrams are given in the first place<br />

of each series of representations in Figure 38 and Figure 41 as well as Figure 44 and<br />

Figure 47. The corange diagrams are resolved to 10% steps of the maximum, while the<br />

associated cotidal behaviour is displayed in steps of 1 /12 of the period. It is obvious from<br />

these delineations that a comparison of several consecutive modes at particular places<br />

in the lake is toilsome and will not help sufficiently for a comparative detailed description<br />

of the different relative local contributions. Therefore, the solutions (13) are represented<br />

additionally by realistic stages of the motion, particularly in time steps of the interface<br />

elevations, as they form in the process of wave propagation. This graphical<br />

resolution has been calculated for the same phase increments as in the cotidal diagrams<br />

and is illustrated by six stages for the first half of the corresponding periods in<br />

Figure 39 and Figure 42 and correspondingly in Figure 45 and Figure 48. The corange<br />

lines in these diagrams are resolved again in 10% steps of the maximum. They are correspondingly<br />

the same in the second half of the oscillation except for the change of the<br />

sign of elevations and need not be repeated.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 61 of 92<br />

Figure 38 : Corange and cotidal diagram of the fundamental internal <strong>seiche</strong> for F = 18.65 (13 April<br />

1989) with the eigen-period of T1 = 407.3 h. The corange lines are given in 10% steps of the maximum,<br />

while the cotidal lines resolve the wave propagation in phase steps of 1 /12 T1 . The oscillation<br />

is cyclonic (anti-clockwise).<br />

If the continuous horizontal representation at different moments is changed to continuous<br />

display of the time history at different sites, a very instructive information results.<br />

This has been calculated for 16 sites on the shore and three in the middle of the lake as<br />

indicated in Figure 37. The resulting 19 diagrams are assembled in Figure 40 and<br />

Figure 43 as complement to the preceding figure and corresponding to the summer<br />

stratification in Figure 46 and Figure 49. The amplitude scale differs in this kind of diagrams<br />

from the previous relative representations, as it is showing the variation normalised<br />

to a maximum of 5 m. Such an assumption meets a realistic size of interface elevations<br />

associated often with internal <strong>seiche</strong>s in Lake Constance. How instructive this<br />

subset of local time histories is, may be perceived from comparable inspection of different<br />

modes for different stratification at a fixed place or at neighbouring places in relation<br />

with the discrete stages, in time, but represented continuously in horizontal dimension<br />

in the preceding figures. These complementary delineations of the amplitude structures<br />

serve a great deal for practical use in the context with the maps of the associated current<br />

field in the surface layer shown in the next section.<br />

A remark on the general structure and its variation with growing influence of the earth’s<br />

rotation on one hand and with gravity on the other hand has to be accomplished, as this<br />

pattern is manifested in the complete series of diagrams in the unpublished German<br />

report and is also apparent in the examples presented here. In the case of high rotation<br />

number, i.e. the case in spring time with extraordinary long eigen-periods, higher amplitudes<br />

are confined to the near-shore region throughout all calculated modes. Moreover,<br />

all amphidromic points in structures up to the 15 th mode are cyclonic and distributed<br />

very regularly, as if suspended on a mid-lake line from one end to the other. This<br />

is true except for two amphidromic points which evade gradually to the southern shore


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 62 of 92<br />

from the tenth mode on, namely towards the mouth of the Bay of Constance and into<br />

the Bay of Rorschach at the greatest width in the eastern half of the basin.<br />

The general regular structure is contrasted by the modes in the late summer stratification.<br />

Here, we realize a reduced confinement of higher amplitudes near the shores in<br />

the wider part of the basin. By contrast, the high amplitudes in the less wide parts as in<br />

the western half and at the eastern end are distributed similar to those of standing<br />

waves at vanishing rotation number. Thus there is a hybrid form of the general structure<br />

which resembles to a certain extent that of standing oscillations.<br />

As to the amphidromic points with increasing modal number, a characteristic difference<br />

appears with the ninth mode, which is represented for spring and summer stratification<br />

in Figure 41 and Figure 47, resp.. Regarding the latter case it has to be premised that<br />

up to the eighth mode all amphidromic systems are cyclonic, their centres aligned along<br />

mid-lake from the western to the eastern end, except for one situated in the Bay of Rorschach.<br />

In the ninth mode shown in Figure 47 a dominant anticyclonic amphidromic<br />

system appears just east of the centre of the lake and a second lateral cyclonic system<br />

appears off the Bay of Constance. For higher modes the structure is seemingly more<br />

complicated, since a few amphidromic systems exist near each other. However, this<br />

pattern develops with increasing mode number more into forms which resemble nodal<br />

lines of standing waves. This is disclosed by crowded cotidal lines connecting certain<br />

adjacent amphidromic points. Such structures represent the change of nodal lines in<br />

standing waves by weak rotational influences in so far, as the jump of the phase by π<br />

across the node is resolved into a steady variation within a narrow band along the node.<br />

Desisting from further considerations of the details an open question has still to be<br />

mentioned. The limitation of the calculations to two typical stratifications with respect to<br />

the structures leaves the user with the uncertainty, whether there are essential variations<br />

of the structures for intermediate cases of stratifications. Certainly, the late summer<br />

situation represents conditions of stratification, which resemble more each other<br />

throughout most of the warm season. Therefore, this part has more bearing on application,<br />

while the displayed case of strong rotational influence in spring may merely provide<br />

rough insight also into weaker autumnal stratification with deepened surface layer.<br />

To the pending completion of the calculations as to other interesting stratifications, the<br />

late summer example allows nevertheless for considerable insight and assessments, if<br />

the assumption of gradual variation of the structures with moderate change of stratification<br />

is correct. As this behaviour can be recognised in the structures with increasing<br />

modal order for both selected stratifications, it is to be expected also for modifications<br />

of these cases.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 63 of 92<br />

Figure 39 : Momentary interface topographies of the first mode shown in Figure 38 at the first six<br />

phase steps of 1 /12 T1 . The lines of zero elevation are congruent with the cotidal lines in Figure 38.<br />

The increment of the corange lines is 10% of the maximum. (13 April 1989, T1 = 407.3 h)


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 64 of 92<br />

Figure 40 : Time histories of interface elevations of the first internal mode as in Figure 38 at 19<br />

sites on the shores and in the lake, however normalised to 5 m maximum amplitude. Map of sites<br />

in Figure 37 (13 April 1989, T1 = 407.3 h)


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 65 of 92<br />

Figure 41 : Corange and cotidal diagram of the ninth internal mode for the stratification on<br />

13 April 1989 with the eigen-period of T9 = 48.2 h (further explanation see Figure 38). The oscillation<br />

around all the amphidromic points is cyclonic (anti-clockwise).


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 66 of 92<br />

Figure 42 : Momentary structures of the amplitude distribution of the ninth mode shown in Figure<br />

41, 13 April 1989, T9 = 48.2 h (further explanation see Figure 39).


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 67 of 92<br />

Figure 43 : Time histories of interface elevations of the ninth internal mode at 19 sites on the<br />

shore and in the lake (13 April 1989, T9 = 48.2 h and further explanation in Figure 40)


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 68 of 92<br />

Figure 44 : Corange and cotidal diagram of the fundamental internal <strong>seiche</strong> for F = 4.16 (30 August<br />

1989) with the eigen-period of T1 = 72.5 h (further explanation see Figure 38). The oscillation is cyclonic<br />

(anti-clockwise).


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 69 of 92<br />

Figure 45 : Momentary interface topographies of the first mode shown in Figure 44 (30 August<br />

1989, T1 = 72.5 h, further explanation see Figure 39)


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 70 of 92<br />

Figure 46 : Time histories of interface elevations of the first mode at 19 sites on the shore and in<br />

the lake (30 August 1989, T1 = 72.5 h, further explanation see Figure 40)


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 71 of 92<br />

Figure 47 : Corange and cotidal diagram of the ninth internal <strong>seiche</strong> for the stratification on 30<br />

August 1989 with F = 4.16 and the eigen-period of T9 = 10.7 h (further explanation see Figure 38).<br />

All the amphidromic points are cyclonic (anti-clockwise) except for the central one in the mid of<br />

the main basin, where the oscillation turns clockwise.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 72 of 92<br />

Figure 48 : Momentary interface topographies of the ninth mode shown in Figure 47 (30 August<br />

1989, T9 = 10.7 h, further explanation in Figure 39


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 73 of 92<br />

Figure 49 : Time histories of interface elevations of the ninth mode at 19 sites on the shore and in<br />

the lake (30 August 1989, T9 = 10.7 h, further explanation in Figure 39)


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 74 of 92<br />

4.6 THE HORIZONTAL STRUCTURES OF THE CURRENTS<br />

The information about the transports associated with the different modes of internal<br />

<strong>seiche</strong>s has important practical reasons as well. It allows for assessment of horizontal<br />

oscillatory displacements and dispersion of either harmful substances or other interesting<br />

compounds in the water, when the site considered is subjected to enhanced activity<br />

of internal <strong>seiche</strong>s. For this purpose, the diagrams have been extended to the horizontal<br />

dependency of the currents. To achieve this, another instructive graphical description<br />

had been elaborated, which meets the vectorial character of this quantity, and delivers<br />

the variation of speed and direction during one wave cycle in synoptic charts. Borrowing<br />

from the illustration of tidal currents, a similar graphical method had been<br />

adopted, namely by showing the current ellipses and, separately, the sense of rotation<br />

indicated at a peculiar moment of the current distribution.<br />

Figure 50 : Current field of the first internal mode in the surface-layer on 13 April 1989 with<br />

T1 = 407.3 h. Upper diagram: current ellipses. The normalised maximum transport at the southern<br />

shore of the entrance of Lake Überlingen is about 600 cm²/s). Lower diagram: momentary current<br />

distribution and sense of current vector rotation.<br />

Before treatment of details, the dimension of transport used here needs a comment.<br />

The numbers of transport are given in cm²/s which results from the vertical integration<br />

of the velocity throughout the upper or lower layer. This quantity is meant as transport<br />

per cm width transverse to the current and thus yields the usual dimension cm³/s.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 75 of 92<br />

The evaluation of the current fields with respect to their part of the eigen-solutions is<br />

unique. However, a clear (and reasonable) graphical representation is only possible for<br />

the relative horizontal distribution of the transports in the upper or lower layer, according<br />

to<br />

V 1 = - V2<br />

→ →<br />

respectively. As the isotherm displacements due to internal oscillations are<br />

much easier to measure than the currents, normally it will be the task to refer the transports<br />

to the amplitudes of the vertical displacements of the interface between the upper<br />

and lower layer, which we have presented in the preceding section. Knowing the phase<br />

velocity and the more easily observable and assessable amplitude of the vertical interface<br />

displacement of a specific mode at any location in the basin it is possible to determine<br />

definite transports by evaluating the relative results of the numerical calculations.<br />

Figure 51 : Current field of the ninth internal mode in the surface-layer on 13 April 1989 with<br />

T9 = 48.2 h. The normalised maximum transport is situated at the mouth of the Bay of Constance<br />

and amounts to about 1800 cm²/s. (further explanation inFigure 50)<br />

To give an example: The vertical displacement of the interface by the fundamental<br />

mode at 30 August 1989 (Figure 44) has its maximum amplitude (100 %) at the very<br />

end of Lake Überlingen. The maximum transport of that mode occurs near the Sill of<br />

Mainau (Figure 53) indicating strong exchange flow between Lake Überlingen and the<br />

main basin of Upper Lake Constance. If we assume the value of 100% to be equivalent<br />

to 100 cm of real interface displacement and normalise the vertically integrated transport<br />

in the upper layer to this quantity, the amplitude of the horizontal transport at the<br />

central position of the Sill of Mainau would be about 4300 cm 2 /s with the definition of<br />

the dimension explained above. Taking the actual depth of the upper and lower layer,<br />

respectively, the horizontal velocities result for the present case to 4300/1500 cm/s and<br />

4300/8500 cm/s as to the lower and upper layer, respectively. These numbers are given<br />

in the legend of Figure 53. How the procedure would work, if the observations came


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 76 of 92<br />

from another location, tells the following example. Take the initial information at the<br />

central position of Lake Überlingen, where amplitudes of about 12 m are of common<br />

occurrence (Bäuerle et al., 1998): From Figure 44 it is inferred that at central Lake<br />

Überlingen the amplitude of the interface displacement is about 90 % of the maximum.<br />

This in turn yields that 13.3 m is the corresponding amplitude at the very end of Lake<br />

Überlingen. Finally, in analogy to the above case of reference, we get<br />

13.3 x 4300/1500 = 38 cm/s and 13.3 x 4300/8500 = 6.7 cm/s as vertically averaged<br />

velocities in the upper and lower layer in the Straits of Mainau, respectively, which correspond<br />

to an amplitude of 12 m measured at a central position of Lake Überlingen.<br />

Figure 52 : Current field of the eleventh internal mode in the surface-layer on 13 April 1989 with<br />

T11 = 39.8 h. The normalised maximum transport occurs at the mouth of the Bay of Constance and<br />

amounts to about 5600 cm²/s. The cross mark near the southern shore in the eastern half of the<br />

lake designates the site of the waste water intake discussed in the text. (further explanation in<br />

Figure 50)<br />

It should be mentioned that the same value of the phase velocity of long internal waves<br />

ci results from different combinations of he and ε according to the relations (1), (2), (8).<br />

Since the eigen-solutions of the problem (5) are uniquely determined for a very phase<br />

speed, ci, the respective definite transport field of the mode in question has to be<br />

evaluated with regard to he, i.e. the associated combination of h1 and h2, which fit together<br />

with ε in the relation (8). In this sense, there is a certain variety of two-layer<br />

manifestations equivalently related to a unique set of eigen-solutions. Thus, for the<br />

same maximum amplitude of a mode of them different definite transports result at a<br />

selected place, just depending on the differences allowed for by both the associated


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 77 of 92<br />

variations of the pair of upper and lower layer depths on one hand and the density difference,<br />

ε, on the other hand. This peculiar variety of two-layer cases covered by one<br />

set of eigenfunctions means a useful advantage, as the same calculations can be exploited<br />

for certain different two-layer approaches.<br />

The current fields of the first and ninth modes are depicted in Figure 50 and Figure 51<br />

for the case in spring and in Figure 53 and Figure 54 for that in late summer. Additionally,<br />

the currents of the eleventh mode are enclosed in Figure 52 and Figure 55 in order<br />

to emphasize the practical use by a peculiar application, which Hollan (1995) carried<br />

out and is outlined below. Except for the reference to the amplitude normalisation the<br />

relative variation of the structures in the current fields can be detected well from the<br />

diagrams. They consist each of a pair, showing the current ellipses in the upper diagram<br />

which indicate the position of the head of the current vector when turning during<br />

one cycle of oscillation. The main axis of the ellipses shows the orientation of the predominant<br />

current during a wave cycle, while the small axis gives the maximum transverse<br />

currents a quarter of the period later than the main currents. In the lower diagram<br />

the information is compiled, which concerns the sense of rotation at a given moment.<br />

Included into this stick representation is the sense of rotation of the current vector. The<br />

illustration of the phase relation and the sense of rotation throughout the lake is secondary<br />

for a rough evaluation, as this information is more important for inspection of<br />

closely adjacent conditions. The assessment of the local intensity of the currents is very<br />

comprehensive in the upper diagram. There is remarkable variation of the intensity and<br />

relative difference between the main and the transverse currents. This is easily perceived<br />

in the different examples and left to the reader for comparison.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 78 of 92<br />

Figure 53 : Current field of the first internal mode in the surface layer on 30 August 1989 with<br />

T1 = 72.5 h. The transport normalised to 1 m maximum vertical displacement of the interface is<br />

4300 cm²/s at the Sill of Mainau, resulting in a current velocity of about 3 cm/s, in the upper layer<br />

(h1 = 15 m) and about 0.5 cm/s in the lower layer (h2 = 85 m), respectively. The evaluation is given<br />

on the previous pages (further explanation in Figure 50).


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 79 of 92<br />

Figure 54: Current field of the ninth internal mode in the surface layer on 30 August 1989, T9 = 10.7<br />

h. The normalised maximum transport (at the mouth of the Bay of Constance) is about 1800 cm 2 /s.<br />

(further explanation in Figure 50)<br />

Not shown here is an example of another additional set of diagrams on current ellipses.<br />

For more detailed local insight, the currents have been depicted in this sort of diagrams<br />

on a larger scale in selected sub-regions. The amount and phase variation with time<br />

along with the sense of rotation has been graphically well resolved. This very instructive<br />

information has been compiled for a few modes in sub-regions covering the complete<br />

Upper Lake Constance. As to this survey report, the quotation of this elaboration may<br />

suffice.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 80 of 92<br />

Figure 55 : Current field of the eleventh internal mode in the surface layer on 30 August 1989,<br />

T11 = 9.6 h. The cross mark near the southern shore in the eastern half of the lake designates the<br />

site of the waste water intake discussed in the text. The normalised maximum transport (at the<br />

mouth of the Bay of Constance) is about 3400 cm 2 /s. The normalised transport at the waste water<br />

intake is about 3200 cm 2 /s. (further explanation in Figure 50)<br />

The above mentioned application by Hollan (1995) elucidates this way of consideration.<br />

For information, the site of a great waste water inlet, which was considered for alternative<br />

construction, is marked by a cross in the lower diagrams of Figure 52 and Figure<br />

55. These figures show the current fields of the eleventh mode in both cases. Compared<br />

to those of the first and ninth mode they differ in the region of this site remarkably,<br />

in particular for the late summer stratification, which is rather representative for the<br />

summer season. The investigation of dispersion conditions in front of this shore section,<br />

which is just aside of the mouth of the Old Rhine, resorted also to the potential activity<br />

of internal <strong>seiche</strong>s in this region. From the comparison of the relative current contributions<br />

by different modes of internal <strong>seiche</strong>s it was deduced, that considerable variability<br />

had to be expected from this process. These conditions can be recognised even from<br />

the examples shown here, as mentioned above. The result of the calculations with this<br />

application is supported by long-time experiences of local fishermen, who reported to<br />

the author (Hollan) about the difficulties with the retrieval of their drift nets in this region<br />

due to transient high vertical current shear in the thermocline. Such phenomenon is<br />

very probably caused by internal <strong>seiche</strong>s, since otherwise vertical shear would be associated<br />

with drift and compensating gradient current during stronger wind attack over the<br />

lake. However, during such conditions fishermen interrupt their work on the lake, what<br />

underscores the first interpretation.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

4.7 REFERENCES<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 81 of 92<br />

Bäuerle E. (1981) : Die Eigenschwingungen abgeschlossener, zweigeschichteter Wasserbecken<br />

bei variabler Bodentopographie. Berichte aus dem Institut für Meereskunde<br />

an der Univ. Kiel, Nr. 85, Kiel, 79pp.<br />

Bäuerle E. and E. Hollan (1983): Calculation of the fundamental and a transverse mode<br />

of internal <strong>seiche</strong>s in Lake Tanganyika. In: Serruya C. and U. Pollingher: Lakes of the<br />

Warm Belt. Appendix, p.499-503, Cambridge, pp.569, Cambridge University Press.<br />

Bäuerle E. (1985): <strong>Internal</strong> free oscillations in the Lake of Geneva. Annales Geophysicae,<br />

Vol. 3, p.199-206.<br />

Bäuerle E. and D. Ollinger (1991): Karten-Dokumentation der internen Seiches des Bodensee-Obersees<br />

für den Gebrauch in der wasserwirtschaftlichen und limnologischen<br />

Anwendung. Unpublished report of the Institut zur Erforschung und zum Schutz der<br />

Gewässer Ottendorf, by contract with the Landesanstalt für Umweltschutz Baden-<br />

Württemberg, Institut für Seenforschung, Langenargen. 10p. with numerous tables and<br />

figures.<br />

Bäuerle E., D. Ollinger and J. Ilmberger (1998): Some meteorological, hydrological, and<br />

hydrodynamical aspects of Upper Lake Constance. In: Bäuerle, E. and Gaedke, U.<br />

(eds.): Lake Constance, Characterization of an ecosystem in transition. Arch. Hydrobiol.<br />

Spec. Issues Advanc. Limnol. 53, p. 31-83.<br />

Hollan E. (1974): Strömungsmessungen im Bodensee. Arbeitsgemeinschaft Wasserwerke<br />

Bodensee-Rhein (AWBR), Sechster Bericht, p.111-187.<br />

Hollan E. and T.J. Simons (1978): Wind-induced Changes of Temperature and Currents<br />

in Lake Constance. Arch. f. Meteorologie, Geophysik u. Bioklimatologie, Ser. A,<br />

Vol 27, p333-373.<br />

Hollan E., D.B. Rao and E. Bäuerle (1980): Free Surface Oscillations in Lake Constance<br />

with an Interpretation of the “Wonder of the Rising Water “ at Konstanz in 1549.<br />

Arch. f. Meteorologie, Geophysik u. Bioklimatologie, Ser. A, Vol. 29, No. 3, p.301-325.<br />

Hollan E. (1995): Ausbreitungsverhalten des Abwassers aus der Kläranlage Altenrhein<br />

im Fernbereich zweier alternativ projektierter Auslauföffnungen an der Halde südwestlich<br />

des Rheinspitz-Canyons im Bodensee-Obersee. <strong>Internal</strong> report of the Institut für<br />

Seenforschung (Landesanstalt für Umweltschutz Baden-Württemberg), by contract with<br />

the Amt für Umweltschutz of the Kanton St. Gallen, Langenargen, pp.17, 1map, 12 figures.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 82 of 92<br />

5 LOCH LOMOND<br />

Although there were no specific modelling activities foreseen within EUROLAKES investigating<br />

<strong>mixing</strong> by internal waves in Loch Lomond it was nevertheless found advantageous<br />

to record recent experiences from field campaigns on the importance of long<br />

internal waves in this lake of irregular shape.<br />

5.1 INTRODUCTION<br />

Loch Lomond is a warm monomictic lake with three very distinct basins. The northern<br />

and central basins are narrow fjord-like (with up to 200 m water depth) and are separated<br />

by a chain of islands from the shallow south basin with a maximum depth of 30 m.<br />

Geologically the loch is a long deep trough of glacial origin and is separated from the<br />

sea by a moraine dam in the south. The mean lake level of Loch Lomond is not higher<br />

than 8 metres above mean sea level. Stratified conditions progressively occur from May<br />

through the summer months particularly in the north basin where a thermocline develops<br />

at a depth of about 15-25 m, separating the warmer epilimnion (ca. 14°C) from the<br />

cooler hypolimnion (ca. 6° C) which was already noted by Slack (1957).<br />

There are, however, only a few past recordings of temperature profiles, the only consistent<br />

approach has been done between 1969 and 1972 by Tippett (1994) with roughly<br />

one measurement profile of temperature and oxygen per month in the centre of every<br />

basin. The result of these experiments are depicted in the following figures for the years<br />

1970 and 1971. Whereas in the south basin thermal stratification occurred only for a<br />

few weeks the deeper portions of the loch show considerable stratification during the<br />

summer months.<br />

Figure 56 : Temperature profiles in the southern basin of Loch Lomond during 1970 and 1971<br />

(data according to Tippett, 1994) with short periods of thermal stratification


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 83 of 92<br />

Figure 57 : Temperature profiles in the northern (above) and central basins (below) of Loch Lomond<br />

during 1970 and 1971 (data according to Tippett, 1994) with pronounced stratification and<br />

high probability of long internal waves


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 84 of 92<br />

5.2 MEASUREMENT INFORMATION ON INTERNAL WAVES<br />

As there is an apparent lack of information concerning the temporal development of the<br />

seasonal thermo- and pycnoclines in Loch Lomond it was decided within the EURO-<br />

LAKES project to carry out a long-term monitoring survey with thermistor-chains<br />

moored at three locations along the main north-south axis of the loch. These thermistor<br />

chains were planned to provide a complete one year data set on the seasonal changes<br />

in water temperature on three vertical profiles from the surface down to a maximum<br />

depth of 50 meters. In conjunction with a number of quasi-synoptic CTD-surveys and<br />

further meteorological monitoring data on solar radiation, air temperature, wind, etc. the<br />

thermistor chain data will provide more comprehensive information to obtain a better<br />

understanding of Loch Lomond’s physics and even more to get a better understanding<br />

of its complete ecosystem.<br />

North Basin<br />

Mid Basin<br />

South Basin<br />

Figure 58 : Location of long-term deployments of thermistor chains in 2002


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 85 of 92<br />

The monitoring survey has been carried out with thermistor chains (AANDERAA Instruments,<br />

Norway) which were provided by the “Institut für Seenforschung” in Langenargen,<br />

Germany within the frame of the scientific co-operation in the research project<br />

EUROLAKES. Management and operations during the survey were carried out by the<br />

project partners from University of Glasgow. The thermistor chain measurements were<br />

planned to cover a time period of at least one year. To gain a high resolution data set<br />

regarding the thermodynamic processes during the development phase of the summer<br />

thermocline a sampling rate of 10 minutes was chosen for the first deployment in the<br />

mid basin in May 2002. Realising that a 10 minutes measuring interval would need data<br />

collection at least every month, it was decided to switch over to a sampling rate of 30<br />

minutes for all deployments from June 2002 onwards. In this report data from the first<br />

measurement period in summer 2002 are used to look at the relative importance of<br />

long internal waves.<br />

Figure 59 : Temperature measured by individual thermistors from May 3 rd to June 8 th 2002 by the<br />

upper chain (in 3 m to 23 m water depth every two metres) in the mid basin (Post, 2002)


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

Depth [m]<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

-5<br />

-10<br />

-15<br />

-20<br />

-25<br />

-30<br />

-35<br />

-40<br />

4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34 36 38 40<br />

Days since 2002/05/01<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Water Temperature in Loch Lomond near Ross Point<br />

Page 86 of 92<br />

Figure 60 : Temperature profiles from 3 rd May to 9 th June 2002 in the mid basin position (Post,<br />

2002)<br />

The recordings from all three locations show considerable short-term fluctuations and at<br />

certain times sudden vertical homogeneity in the thermal conditions. Interpretation of<br />

these (horizontally seen) point measurements is not as straightforward as it might seem<br />

because Loch Lomond has a very irregular shape. This means that a generalisation of<br />

these profiles as “basin characteristic” is not possible. Investigations with 3D models<br />

(described in other reports of EUROLAKES) did show strong wind-driven currents with<br />

pronounced upwelling/downwelling processes near steep lake shores and the possible<br />

occurrence of longer period internal standing waves in the fjord-like section of the lake.<br />

Generally, however, it can be stated that long internal waves in the south basin are very<br />

intermittent and associated <strong>mixing</strong> processes will be much smaller than the turbulent<br />

<strong>mixing</strong> associated with the wind stress at the surface. Therefore in this report analysis<br />

of data is confined to the mid and north basin locations where stratification is strong in<br />

the upper 40 metres of the water column. Measurements are scheduled to proceed until<br />

spring 2003 but we are concentrating here on the summer situation.<br />

12.7<br />

12.2<br />

11.7<br />

11.2<br />

10.7<br />

10.2<br />

9.7<br />

9.2<br />

8.7<br />

8.2<br />

7.7<br />

7.2<br />

6.7


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 87 of 92<br />

5.3 ANALYSIS AND INTERPRETATION<br />

In order to obtain a clearer picture of the kind of temperature fluctuations occurring in<br />

the upper water column in Loch Lomond the thermistor recordings from both north and<br />

mid basin locations were analysed statistically for the period of 8 th August and 18 th<br />

September 2002.<br />

In the following temperature histograms are shown for two vertical levels (6 metres and<br />

45 metres below surface) for north and mid basin deployments depicting the existence<br />

of pronounced long internal waves in Loch Lomond with a stronger variability of temperature<br />

in the central basin.<br />

Frequency analysis provides a few conspicuous periods ranging between about 5 hours<br />

and 24 hours (a frequency of 0.25 in the figures is connected with a period of 2 hours =<br />

time step of recording divided by frequency). The periods coincide with theoretical values<br />

for long standing waves in a narrow channel of 20 km length (north + mid basin) for<br />

strong vertical density stratification near the surface and less strong ones below 20 metres.<br />

The results for the mid basin prove quite clearly that it reacts as an appendix to the<br />

north basin because the long periods cannot be explained by its own basin length of<br />

roughly five kilometres. The differences in periods are due to different thermal stratification<br />

conditions.<br />

5.4 REFERENCES<br />

Post, J. (2002): Thermistor measurements in Loch Lomond. HYDROMOD Scientific<br />

Consulting, Wedel – unpublished report.<br />

Slack H. D. (ed.) (1957) Studies on Loch Lomond 1. Blackie and Son Ltd., Glasgow.<br />

Tippett R. (1994): An Introduction to Loch Lomond, Hydrobiologia, 290, 11-15. Kluwer<br />

Academic Publishers, Dordrecht / Boston / London.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

14.0 14.4 14.8 15.2 15.6 16.0 16.4 16.8 17.2 17.6 18.0 18.4 18.8<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 88 of 92<br />

Figure 61 : Histogram of temperature at 6 metres water depth for north basin (above) and mid basin<br />

(below) during 8 th August to 18 th 0.0<br />

15.0 15.4 15.8 16.2 16.6 17.0 17.4 17.8 18.2 18.6 19.0<br />

September 2002


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

5<br />

4<br />

3<br />

2<br />

1<br />

6.2 6.4 6.6 6.8 7.0 7.2 7.4 7.6 7.8 8.0 8.2 8.4<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 89 of 92<br />

Figure 62 : Histogram of temperature at 45 metres water depth for north basin (above) and mid basin<br />

(below) between 8 th August and 18 th 0<br />

7.4 7.6 7.8 8.0 8.2 8.4 8.6 8.8 9.0 9.2<br />

September 2002


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

8 spctr.h41.3.detr.MB<br />

spctr.h31.3.detr<br />

spctr.h.20.3.detr<br />

spctr.h16.3.detr<br />

6<br />

4<br />

2<br />

0<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

0.00 0.05 0.10 0.15 0.20 0.25<br />

frequency<br />

0.00 0.05 0.10 0.15 0.20 0.25<br />

frequency<br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

spctr.h50.95.detr<br />

spctr.h45.95.detr<br />

spctr.h40.95.detr.NB<br />

spctr.h35.95.detr<br />

spctr.h30.95.detr<br />

spctr.h25.95.detr<br />

spctr.h20.95.detr<br />

Page 90 of 92<br />

Figure 63 : Frequency analysis of temperature fluctuations in mid basin (at 16 to 41 metres, picture<br />

above) and north basin (20 to 51 metres, picture below) for low pass-filtered data.


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 91 of 92<br />

6 CONCLUSIONS<br />

The better understanding and knowledge of the internal gravity waves (<strong>seiche</strong>s) in deep<br />

large lakes were the main objectives of this <strong>study</strong>. We have focused the work principally<br />

on the determination of the waves properties through measurements analysis and numerical<br />

simulation. The <strong>seiche</strong>s qualification has been done all over the four studied<br />

lakes. The influence of the <strong>seiche</strong>s on <strong>mixing</strong> processes was also studied.<br />

Several approaches of the internal gravity wave in large deep lakes have been presented:<br />

• Analysis of existing measurement on Loch Lomond, Lac du Bourget, Lac<br />

Léman,<br />

• Linear analysis (eigen-value method): Lake Constance,<br />

• Numerical 3D model analysis: Lac du Bourget, Lac Léman<br />

6.1 LAC DU BOURGET<br />

The analysed measurements are shown the existence of internal gravity wave in the<br />

Lac du Bourget with a period between 40hours to 80 hours depending on the vertical<br />

stratification structure.<br />

The simulation using the TELEMAC-3D model is used to reproduce one <strong>seiche</strong> event<br />

on the Lac du Bourget. The numerical results are in a quite good qualitative agreement<br />

with the measurements.<br />

The numerical results are given some interesting new information about the hydraulic<br />

respond of the lake under wind forcing. After the generation of the internal <strong>seiche</strong>s, the<br />

wave propagation is showing a rotating structure, due to the Coriolis influence, with an<br />

amphidromic point at the middle of the lake (Figure 16).<br />

6.2 LAC LÉMAN<br />

An analysis of internal <strong>seiche</strong>s dynamics was carried out for Lac Léman combining field<br />

measurements and numerical modeling. Using field data of temperature, currents and<br />

surface elevation, it has been shown that only two modes of internals <strong>seiche</strong>s are sufficiently<br />

excited in Lac Léman to be considered significant. The first one is a Kelvin wave<br />

and the second one is a Poincaré wave. Model calculations have indicated that other<br />

<strong>seiche</strong> modes can only be excited by winds from certain directions. However, due to the<br />

topographic constraints particularly in the eastern part of the lake basin the wind field<br />

over the lake is strongly canalized and these winds do not exist in nature.<br />

From our analysis of field studies of the longterm mean conditions of <strong>mixing</strong> it is indicated<br />

that internal <strong>seiche</strong>s are important in providing vertical <strong>mixing</strong>. Recently, it has<br />

been shown though that most of this <strong>mixing</strong> is actually generated in the near shore<br />

zone and then propagates into the open waters (Wuest et al., 2000). Thus, the interaction<br />

between near shore zones and the open water is also important for <strong>mixing</strong>. Furthermore,<br />

we have pointed to the importance of the interaction with the sloping sides of<br />

the lake and short progressive internal waves (Thorpe and Lemmin, 1999a, Lemmin et


Integrated Water Resource Management for Important Deep European Lakes and their Catchment Areas<br />

EUROLAKES<br />

<strong>D28</strong>: <strong>Internal</strong> <strong>seiche</strong> <strong>mixing</strong> <strong>study</strong><br />

FP5_Contract No.: EVK1-CT1999-00004<br />

Version: 1.2<br />

Date: 24.08.2004<br />

File: <strong>D28</strong>.doc<br />

Page 92 of 92<br />

al., 1998). These waves and their breaking play a role in the production and redistribution<br />

of currents and stratification as well as <strong>mixing</strong> (Thorpe and Jiang, 1998). From our<br />

studies it appears that short progressive internal waves are often produced in the passage<br />

of non-linear internal <strong>seiche</strong>s (Thorpe et al., 1996).<br />

The measured characteristics of the velocity time series (namely the non-linear wave) in<br />

the case of a strong wind event could be reproduced. In addition, the good agreement<br />

of the measured and computed structure of the temperature profile suggest that the<br />

turbulence model yields reasonable turbulent diffusivities.<br />

6.3 LAKE CONSTANCE (BODENSEE)<br />

The method of calculating the eigen-periods of free internal oscillations in a two-layermodel<br />

also permits a relatively simple evaluation of the influence of the stratification on<br />

the eigen-periods. But also the variation of the structure with respect to the stratification<br />

is calculated at the same time and provides the characteristical differences, which result<br />

from strong or diminishing influence of the earth´s rotation via varying stratification.<br />

The application of the eigen-value method should be calculated for other lakes in order<br />

permit the comparison and evaluation of the lake-specific characteristics<br />

6.4 LOCH LOMOND<br />

In order to obtain a clearer picture of the kind of temperature fluctuations occurring in<br />

the upper water column in Loch Lomond the thermistor recordings from both north and<br />

mid basin locations were analysed statistically for the period of 8 th August and 18 th<br />

September 2002. Frequency analysis provides a few conspicuous periods ranging between<br />

about 5 hours and 24 hours.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!