196
Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng. (Hons.) (Civil Eng.) M.Sc. (Water Resour. Eng.) Facultad de Minas, Universidad Nacional de Colombia, Medellín, Colombia This thesis is presented for the degree of Doctor of Philosophy of The University of Western Australia Centre for Water Research January 2007

Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Observations of energy transfer mechanisms associated with internal

waves

Evelio Andrés Gómez Giraldo

B.Eng. (Hons.) (Civil Eng.) M.Sc. (Water Resour. Eng.)

Facultad de Minas, Universidad Nacional de Colombia, Medellín, Colombia

This thesis is presented for the degree of Doctor of Philosophy of

The University of Western Australia

Centre for Water Research January 2007

Page 2: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

ii

Page 3: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

iii

Abstract

Internal waves redistribute energy and momentum in stratified lakes and

constitute the path through which the energy that is introduced at the lake scale is

cascaded down to the turbulent scales where mixing and dissipation take place.

This research, based on intensive field data complemented with numerical

simulations, covers several aspects of the energy flux path ranging from basin-

scale waves with periods of several hours to high frequency waves with periods of

few minutes.

It was found that, at the basin-scale level, the horizontal shape of the lake

at the level of the metalimnion controls the period and modal structure of the

basin-scale natural modes, conforming to the dispersion relationship of internal

waves in circular basins. The sloping bottom, in turn, produces local

intensification of the wave motion due to focusing of internal wave rays over

near-critical slopes, providing hot spots for the degeneration of the basin-scale

waves due to shear instabilities, nonlinear processes and dissipation.

Different types of high-frequency phenomena were observed in a stratified

lake under different forcing conditions. The identification of the generation

mechanisms revealed how these waves extract energy from the mean flow and the

basin-scale waves. The changes to the stratification show that such waves

contribute to mixing in different ways. Consequently, each type of wave

constitutes a different energy flux path from large-scale to small-scale motions to

mixing in stratified lakes. First, periodic high frequency waves are generated by

shear instabilities in the surface layer of the lake during strong wind periods.

These waves produce active mixing in the surface layer but do not contribute

Page 4: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

iv

significantly to the horizontal transport of mass and energy because they dissipate

before being able to propagate any distance. Second, periodic high frequency

waves are also generated over topographic features that enhance the basin-scale

wave induced shear. They contribute to localized mixing in the metalimnion and

dissipate quickly. Third, internal solitary-like waves are generated by nonlinear

processes at the crest of steepened basin-scale waves. These waves propagate

large distances without generating significant mixing in the interior of the lake,

and lose most of their energy when reflected at the sloping end of the lake, leading

to mixing in the benthic boundary layer. Last, fronts with strong surface

horizontal density gradient generated by full upwelling events develop into

internal bores. According to its strength, a bore behaves as a gravity current, a

non-breaking undular or a breaking undular bore, and may contribute actively to

vertical mixing as it propagates.

Detailed field observations were used to develop a comprehensive

description of an undocumented energy flux mechanism in which shear-

instabilities with significant amplitudes away from the generation level are

produced in the surface layer due to the shear generated by the wind. The vertical

structure of these instabilities is such that the growing wave-related fluctuations

strain the density field in the metalimnion triggering secondary instabilities. These

instabilities also transport energy vertically to the thermocline where they transfer

energy back to the mean flow through interaction with the background shear.

Page 5: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

v

Contents

Abstract iii

List of Figures ix

List of Tables xvii

Dedication xix

Acknowledgements xxi

Preface xxiii

1 Introduction 1

1.1 Motivation 1

1.2 Overview 4

2 Spatial Structure of the dominant basin-scale internal waves in

Lake Kinneret 7

2.1 Abstract 7

2.2 Introduction 8

2.3 Field site 12

2.4 Field equipment 12

2.5 Observed wind pattern 13

2.6 Measured basin scale internal waves activity 13

2.6.1 Propagation 14

2.6.2 Azimuthal structure 15

2.6.3 Radial structure 16

2.6.4 Vertical structure 16

2.7 Numerical simulations 17

Page 6: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

vi

2.7.1 Validation 18

2.7.2 Numerical damping in ELCOM 20

2.7.3 Basin with flat bottom and irregular horizontal shape 23

2.7.4 Lake bathymetry with variable depth 24

2.8 Interaction of the wind and the dominant natural modes 26

2.8.1 Reset of the phase and modification of the period 26

2.8.2 Resonance 27

2.8.3 Spatial structure of the observed oscillations 28

2.9 Discussion 30

2.10 Conclusions 36

2.11 Acknowledgements 37

3 Classes of high-frequency motions and their implications for

mixing: observations from Lake Constance 61

3.1 Abstract 61

3.2 Introduction 62

3.3 Field site 65

3.4 Field equipment 65

3.5 Methods 66

3.6 Background stratification, forcing events, and basin-scale motions 68

3.7 High frequency waves. Prominent features 69

3.8 Waves with periods between 4 and 13 minutes 70

3.9 High frequency event I 72

3.10 High frequency event II 75

3.11 Discussion 80

Page 7: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

vii

3.11.1 High-frequency waves generated by shear instability 80

3.11.2 Solitary waves 81

3.11.3 Internal bore 82

3.12 Conclusions 86

3.13 Acknowledgments 87

Appendix. Derivation of U 87

4 Wind-shear Generated High-Frequency Internal Waves as

Precursors to Mixing in Lake Kinneret 107

4.1 Abstract 107

4.2 Introduction 108

4.3 Field site 111

4.4 Field equipment 112

4.5 Analysis methods: Linear stability analysis 112

4.6 Observed background conditions 114

4.7 Characteristics of the observed high frequency waves 116

4.8 Results of the stability analysis 118

4.9 Perturbation kinetic energy 122

4.10 Discussion 123

4.11 Conclusions 128

4.12 Acknowledgments 129

5 Conclusions 149

5.1 Summary 149

5.2 Future work 151

Bibliography 153

Page 8: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

viii

Page 9: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

ix

List of Figures

2.1 Bathymetry of Lake Kinneret. 41

2.2 (a) Wind speed and (b) wind direction (meteorological convention)

measured 1.5 m over the lake surface at station Tv. (c) Power

spectral density of the wind speed in (a). 42

2.3 Record of isotherm displacements. 43

2.4 Power spectra of measured and simulated vertical displacement of

the 23°C isotherm. 44

2.5 (a) Mean temperature profile at station T4 for days 170.5 to 171.5

in 2001. (b) Associated buoyancy frequency. 45

2.6 (a) Depth of the 23°C isotherm displacements at stations along the

20 m isobath (with 10 m offset). Vertical displacement band-

passed around (b) 24 hours (-10 m offset), and (c) 12 hours (-5 m

offset). 46

2.7 Temporal evolution of the 24-h component of the 23°C isotherm

vertical displacement along the (a) azimuthal and (b) radial

direction, and of the 12-h period component along the (c)

azimuthal and (d) radial direction. 47

2.8 Profiles of coherence squared (solid line for field data and dashed

line for ELCOM results), and phase (circles for field data and

triangles for ELCOM results) of the vertical displacement of 24-h

(a, b, and c) and 12-h (d, e, and f) period oscillations. 48

2.9 (a-f) Depth of the 23°C isotherm from field data and ELCOM

Page 10: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

x

simulation; (g-l) 23°C isotherm vertical displacements band-pass

filtered around 24 h; (m-r) vertical displacements band-pass

filtered around 12 h.

49

2.10 Dominant natural modes of a basin with the horizontal shape of

lake Kinneret 15 m isobath, flat bottom, and continuous

stratification. (a) Power spectra of isotherm displacement and (b)

cyclonic component of the rotary power spectra of isopycnal

velocity for the 23-h-period natural mode; (c) power spectra of

isotherm displacement and (d) difference between the anticyclonic

and cyclonic components of the rotary power spectra of isopycnal

velocity for the 10.5-h-period natural mode. 50

2.11 Dominant natural modes of lake Kinneret. (a) Power spectra of

isotherm displacement and (b) cyclonic component of the rotary

power spectra of isopycnal velocity for the 23-h-period natural

mode; (c) power spectra of isotherm displacement and (d)

difference between the anticyclonic and cyclonic components of

the rotary power spectra of isopycnal velocity for the 10.5-h-period

natural mode. 51

2.12 Snapshots of the relevant component of the 10.5 h band-passed

velocity along the north-south axis of Lake Kinneret. (a) and (c)

show velocity perpendicular to the transect with positive values

indicating velocities towards the east; (b) and (d), velocity along

the transect with positive values indicating velocities towards the

south. 52

Page 11: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xi

2.13 (a) West-east component of the wind at Tv and depth of the 23°C

at stations (b) Tf and (c) T7. 53

2.14 (a) Momentum input of the daily wind events, (b) temporal

location of the daily wind events, (c) elapsed time between the

ends of two consecutive daily wind events (the dashed line

indicates Δt = 22.6 h), (d) amplitude of the Kelvin wave, and (e)

time of arrival of the Kelvin wave crest. 54

2.15 Lake Kinneret with measured forcing. (a) Power spectra of

isotherm displacement and (b) cyclonic component of the rotary

power spectra of isopycnal velocity at 24 h; c) power spectra of

isotherm displacement and d) difference between the anticyclonic

and cyclonic components of the rotary power spectra of isopycnal

velocity at 12 h. 55

2.16 RMS vertical displacement profile at station Tv calculated from

measured and simulated temperature. 56

2.17 Snapshots of the 21°C isothermal velocity for the simulation of the

flat bottom basin with the horizontal shape of the 15 m isobath of

lake Kinneret and continuous stratification after an initial tilt. 57

2.18 (a) Dispersion relationships of the two cells of the 10.5-h-period

natural mode of a basin with the shape of lake Kinneret at the level

of the thermocline, superimposed in the dispersion relationships

derived by Antenucci and Imberger (2001). (b) Dependence of the

ratio of radii of the anticyclonic and cyclonic cells on the Burger

number of the anticyclonic cell. 58

Page 12: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xii

2.19 Shaded areas indicate where the ratio of the bottom and critical

slopes is between 1/1.5 and 1.5 for (a) the 22.6-h period Kelvin

natural mode and (b) the 10.5-h period natural mode. 59

3.1 Lake Constance (47°39’N, 9°18’E) bathymetry and location of

Lake Diagnostic Systems with thermistor chains and wind sensors. 91

3.2 1-h averages of (a) wind speed and direction, (b) isotherm

contours, and (c) buoyancy frequency at s7. 92

3.3 (a) 10-min average of wind speed (solid line) and direction (dots)

at s2, (b) 10-s time series of integrated potential energy at s2, and

(c) wavelet power spectra of integrated potential energy at s2. 93

3.4 Isotherm displacement showing high frequency internal waves (a)

during strong wind at s5 and (b) during the passage of the Kelvin

wave trough at s3. 94

3.5 Wind speed, integrated potential energy and log2 of wavelet power

of integrated potential energy in the 4- to 13-min band for some of

the stations 304 and 305.5. 95

3.6 Wind speed, integrated potential energy and log2 of wavelet power

of integrated potential energy in the 4- to 13-min band for some of

the stations between days 299 and 300.5. 96

3.7 (a) Depth of the 9°C isotherm from field data and ELCOM

simulation at s3 located over the sill of Mainau. (b) Regions in the

time-depth plane with N larger than 0.002 Hz where the scaled Rig

was below the critical value of 0.25. 97

3.8 0.5°C interval isotherm contours at s3 showing the fission of a

Page 13: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xiii

solitary-like wave. 98

3.9 Wind speed, integrated potential energy and log2 of wavelet power

of integrated potential energy in the 13-min to 1.2-h band for the

stations in lake Überlingen and western end of the main basin. 99

3.10 Isotherm displacements generated by the solitary waves and the

wavelet power in the 13-min to 1.2-h band and in the 4- to 13-min

band. (a), (b), and (c) show the incident solitary waves from s3 to

s1, and (d), (e), and (f) show the reflected solitary waves from s1 to

s3. 100

3.11 Wind speed, integrated potential energy and log2 of wavelet power

of integrated potential energy in the 13-min to 1.2-h band for

stations s5 and s7. 101

3.12 Evolution of surface water temperature and velocity from ELCOM

simulations. Thick arrows show the wind speed. 102

3.13 0.5°C interval isotherm contours showing the propagation of the

warm front through (a) s5, (b) s3, (c) s2, and (d) s1. (e) shows the

front reflected from the sill as it passes s5. Shaded regions are

magnified in Fig. 3.14. 103

3.14 Magnified views of shaded areas in Fig. 3.13. 104

3.15 Isotherms displacement showing an undular internal bore as it

passes s6. (a) Incident bore, (b) reflected bore. 105

4.1 Map of Lake Kinneret. The filled circle at B indicates the location

of the 5-LDS array. The inset sketches the relative location of the

LDSs (filled circles), with distances and direction of the lines

Page 14: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xiv

specified in Table 4.1. The cross near the centre of the large

triangle indicates the approximate location of the PFP casts.

134

4.2 Wind and temperature observations at TB1. (a) Ten-minute

average wind speed and direction. (b) Temperature contours at 2°C

intervals; the bottom isotherm is 17°C. (c) and (d) Magnified views

of shaded regions c and d in (b). 135

4.3 Power spectral density of 23°C isotherm vertical displacements at

station TB1 for the entire length of the record. 136

4.4 (a) Regions with Richardson number smaller and larger than 0.25

are coloured in black and grey respectively. The white lines are

isotherms at 2°C intervals. (b) Direction of the dominant shear 137

4.5 (a) 23°C isotherm vertical displacements band-pass filtered around

0.0056 Hz for the entire record period. (b) and (c) Magnification

for two periods of strong high frequency internal wave activity. 138

4.6 Power spectra of the 23°C isotherm displacements for periods

1500, 1630 and 1900, and of the27°C isotherm displacement for

period 1500. 139

4.7 Selected vertical velocity from PFP data for period (a) 1500, (b)

1630 and (c) 1900. 140

4.8 Mean background profiles for the three periods were stability was

investigated during day 179. (a) Density, (b) and (c) meridional

(U0) and zonal (U90) components of velocity. (d) Shading indicates

the direction of shear for period 1500, at those depths where Rig <

0.25. (e) and (f) Same as (d) for periods 1630 and 1900

Page 15: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xv

respectively. 141

4.9 (a) Growth rate of unstable modes for direction 101.25° during

period 1500. (b), (c), (d), and (e) Vertical structure of the modes I,

II, III, and IV. 142

4.10 Maximum growth rate of unstable modes in horizontal

wavenumber space. for (a) period 1500, (b) period 1630, and (c)

period 1900. 143

4.11 (a) Growth rate of unstable modes type A for direction 112.5°

during period 1630. (b), (c), (d), and (e) Vertical structure of the

modes I, II, III, and IV. 144

4.12 (a) Growth rate of unstable modes type A for direction 101.25°

during period 1900. (b), (c), (d), and (e) Vertical structure of the

modes I, II, III, and IV. 145

4.13 Terms in the kinetic energy equation for the unstable mode

responsible for the high frequency oscillations during period 1900.

(a) Complex vertical velocity. (b) Shear production of perturbation

kinetic energy, (c) work done by gravity, and (d) vertical flux. 146

4.14 (a) Vertical velocity due to the unstable mode IV in Fig. 12e after

it has grown from an initial (arbitrary) maximum amplitude of 0.02

m s-1 at 4 m depth during 1000 s until it reached a maximum

amplitude of 0.05 m s-1. (b) Associated density perturbations. (c)

Total density obtained by adding the background density in Fig

4.8a and the density perturbation in Fig. 4.14b. 147

Page 16: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xvi

Page 17: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xvii

List of Tables

2.1 Details of the deployment 38

2.2 Results of coherence and phase for 23°C isotherm vertical

displacement for field data and ELCOM results. 39

2.3 Estimates of dissipation time scale, damped natural period, and

resonant amplitude, obtained from the field data and from

simulations with different grid size. 40

3.1 Characteristics of the internal bore generated after the full

upwelling event. 90

4.1 Characteristics of the configuration of the LDS array. 130

4.2 Periods of analysis of the characteristics of the high frequency

internal waves. 131

4.3 Characteristics of measured waves. θ is the direction of

propagation. 132

4.4 Propagation characteristics of the fastest growing modes of the

three periods considered. 133

Page 18: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xviii

Page 19: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xix

Dedication

This work is dedicated to Natalia, my very special sister.

Page 20: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xx

Page 21: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xxi

Acknowledgements

I thank Professor Jörg Imberger and Dr. Jason Antenucci for their

supervision of this work. Both provided invaluable guidance, encouragement,

support and example.

I want to acknowledge the Field Operations Group at the Centre for Water

Research, the Kinneret Limnological Laboratory, and the Institut für Wasserbau at

the University of Stuttgart, for planning and running successfully the field

experiments in Lake Kinneret and Lake Constance. The personal effort of Sheree

Feaver, Roger Head, Carol Lam, Greg Attwater, Jochen Appt, Simone Moedinger,

Peter Yeates, Penny VanReenen and Jörg Imberger is highly appreciated. I am

also grateful to the Israel Water Commission and the Deutsche

Forschungsgemeinschaft for funding the research projects.

My stay at the University of Western Australia was supported by an

International Postgraduate Research Scholarship (IPRS), a UWA Scholarship for

International Research Fees (SIRF), and an ad-hoc CWR scholarship.

I am grateful for the constructive and inspirational scientific discussions

and comments from Geoff Wake, Roman Stocker, Leon Boegman, Arthur

Simanjuntak, and Erich Bäuerle. Thanks to Patricia Okely for proofreading some

parts of this thesis and to Carlos Ocampo for printing this final version.

I want to thank all the staff, visitors and students of the Centre for Water

Research for making this a very nice place to do research. I want to thank Ryan,

Geoff, Seba, Clelia, Carlos, Sam, Peter, Vadim, Leon, Arthur, Sheree, Roman,

Daniel, Alan, Ingrid and Yanti for their friendship and support during hard times.

Page 22: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xxii

Finally, a very special thanks to Mónica for her love, support, cooperation

and enormous patience.

Page 23: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xxiii

Preface

The main body of this thesis (Chapters 2-4) is a compilation of three self-

contained papers written for journal publication. Each of these papers contains a

review of the relevant literature. The introductory Chapter 1 places the individual

papers in the limnological context and, to avoid unnecessary duplication, the

literature review is orientated to identify the relevance of this work. Chapter 5

draws together the conclusions from the individual papers and makes suggestions

for future work.

Chapter 2 has been published by the Journal Limnology and

Oceanography as: Gómez-Giraldo, A., J. Imberger, and J. P. Antenucci. 2006.

Spatial structure of the dominant basin-scale internal waves in Lake Kinneret.

Limnol. Oceanogr. 51: 229-246.

Chapter 3 will be submitted shortly for publication at the Journal

Limnology and Oceanography. The manuscript is jointly authored with J.

Imberger and J. P. Antenucci.

Chapter 4 has been submitted for publication at the Journal Limnology and

Oceanography as: “Wind-shear generated high-frequency internal waves as

precursors to mixing in a stratified lake”, by A. Gómez-Giraldo, J. Imberger, J. P.

Antenucci, and P. S. Yeates.

Although inspired and directed by my supervisors Prof. Jörg Imberger and

Dr. J. P. Antenucci, the work contained in this thesis is wholly my own.

Page 24: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

xxiv

Page 25: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

1

Chapter 1

Introduction

1.1. Motivation

Lakes are often thermally stratified due to the exchange of thermal energy

with the atmosphere at the surface. The stirring generated by the wind produces

mixing at the surface creating a homogeneous surface layer (the epilimnion). The

colder water in the bottom of the lake (the hypolimnion) is also relatively

homogeneous and is separated from the epilimnion by a thin layer with a strong

density gradient (the metalimnion). The strong buoyancy in the metalimnion

restricts vertical movement and this layer acts as a physical barrier that inhibits

mixing between the surface and bottom waters, playing an important role in lake

ecology. Primary production takes place in the epilimnion due to the presence of

light and oxygen constantly supplied through the interaction with the atmosphere.

However, the supply of nutrients to sustain primary production can be insufficient

when the river inflows are reduced. In contrast, the hypolimnion is nutrient rich

due to the breakdown of organic matter, but lacks light and oxygen. The regular

supply of nutrients that is required to sustain primary productivity in the

epilimnion can only come from mixing with hypolimnetic water if the surface

inflows are significantly reduced, as is the case for most lakes in medium latitudes

during summer (Ostrovsky et al. 1996; Wetzel 2001).

To comprehend lake ecology it is then necessary to construct a clear

picture of the distribution of the buoyancy flux (usually quantified indirectly by

the associated dissipation of turbulent kinetic energy) and the path along which

the energy becomes available at the turbulent scale (Imberger 1998). Energy is

Page 26: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

2 Chapter 1

introduced in the lake by the action of wind blowing over the lake surface. In

addition to mixing the surface layer, the wind stress piles up the surface water at

the downwind end of the lake, generating a longitudinal pressure gradient that, in

turn, produces an opposite tilt of the metalimnion with upwelling of the

hypolimnion at the upwind end of the lake (Mortimer 1974).

When the wind stops, the relaxation of the longitudinal pressure gradient

generated by the tilt of the metalimnion sets up periodic motions with

characteristic lengths and periods defined by the stratification, the shape of the

basin and, when the effects of the Earth’s rotation has an important effect, the

Coriolis force (Imboden 1990). These are basin-scale internal waves, of which

there are two types. The first are gravity waves, governed by a balance between

gravity, Coriolis and inertia forces (Mortimer 1974). These waves dominate the

spectra of isotherm displacement in stratified lakes and are widely documented

(Lemmin 1987; Antenucci et al. 2000; Rueda et al. 2003). If the internal motions

in the lake are unaffected by the Earth’s rotation, these waves are standing

seiches; otherwise, they are of the Kelvin and Poincaré types. The second type of

internal basin-scale waves are topographic waves, which are governed by the

conservation of potential vorticity (Csanady 1976; Stocker and Hutter 1987).

Basin-scale waves generate mixing along the bottom of the lake by

sloshing (Taylor 1993) and by shear instabilities (Lemckert and Imberger 1998),

although this only leads to exchange of epilimnetic and hypolimnetic masses of

water at the locations where the metalimnion intercepts the bottom of the lake. To

determine such locations, how often the metalimnion sweeps them and the relative

strength of the wave-related currents, the periodicity and structure of the basin-

scale internal waves must be known. The structure of the basin-scale internal

Page 27: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Introduction 3

waves also creates heterogeneity in processes such as resuspension of sediments

(Fricker and Nepf 2000; Marti 2004) and the associated increase in productivity

(Ostrovsky et al 1996; MacIntyre et al 1999). Analytical models with simple

geometries that consider a layered stratification (Heaps and Ramsbottom 1966;

Csanady 1967; Antenucci and Imberger 2001a) or a continuous vertical density

profile (Csanady 1972; Monismith 1987) have been developed. They produce

good estimates of the periods and the basic features of the waves, but are unable

to reproduce any effects of an irregular basin shape and a sloping bottom.

Numerical eigenvalue models have been used improve the estimates of the spatial

structure of the internal basin-scale waves (Schwab 1977; Horn et al. 1986;

Münnich 1996; Bäuerle 1998; Fricker and Nepf 2000) but none of them can

consider the combined effect of irregular shape, continuous stratification, and

Coriolis force.

Smaller scale internal waves with periods of minutes are also ubiquitous in

stratified lakes (Mortimer et al. 1968; Hamblin 1977; Thorpe et al 1996; Saggio

and Imberger 1998; Antenucci and Imberger. 2001b) and are generated by

phenomena that can be summarized as shear instabilities or nonlinear processes

(Boegman et al. 2003). In some cases, the high-frequency waves get their energy

directly from external forcing (Hamblin 1977; Farmer 1978; Antenucci and

Imberger 2001b) but several field observations indicate that very often they

extract energy from the basin-scale internal waves (Thorpe et al. 1996; Saggio and

Imberger 1998; Stevens 1999). High-frequency internal waves generated by

shear-instability are believed to propagate over short distances before they loss

their energy, while those generated by nonlinear processes are believed to travel

long distances without losing a significant amount of energy until they reach the

Page 28: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

4 Chapter 1

sides of the lake and break (Boegman et al. 2003), contributing to mixing in the

benthic boundary layer. This points to a different participation of the different

types of high frequency waves in the energy flux path in stratified lakes with

shear-generated waves transferring energy from basin-scale motions to turbulence

locally, and with solitary-like waves extracting energy from the basin-scale waves

in one location and transferring it to turbulence in a remote location.

The books by Drazin and Reid (1981) and Drazin and Johnson (1989)

provide a good summary of the analytical theory of shear-instabilities and

nonlinear evolution of waves, and the behaviour of high-frequency waves under

controlled laboratory or numerical experiments is well known (Thorpe 1971;

Hazel 1972; Horn et al. 2001; Vlasenko and Hutter 2002; Boegman et al 2005a).

In the field, however, the description of the high-frequency waves is normally

limited to some characteristics of the waves due to the lack of high spatial and

temporal data resolution.

1.2. Overview

This thesis seeks to contribute to the understanding of the energy flux path

in stratified lakes from basin-scale internal waves, through high frequency internal

waves, to turbulence. It does not concentrate on quantification of energy fluxes,

but rather on identifying the mechanisms responsible for them. The approach

followed relies strongly on field measurements of temperature collected at a high-

frequency rate and simultaneously at several locations in the lake. The

information extracted from the field data is complemented with numerical

simulations to improve the spatial resolution.

Page 29: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Introduction 5

Here, the vertical structure of all the types of internal waves considered is

described in terms of modes in a continuously stratified medium, unless the

description in terms of vertically propagating rays is specifically invoked.

Chapter 2 investigates the influence of the bathymetry on the structure of

basin-scale internal waves considering continuous stratification, Coriolis force

and the real bathymetry of the lake. Two different effects of the bathymetry are

studied: the effect of the horizontal shape and the effect of the sloping bottom.

Chapter 3 describes observations made at one lake of several types of high

frequency internal waves responding to different forcing and background

conditions. The generation mechanisms of the high-frequency internal waves are

identified as well as how the different waves contribute in different ways to the

heterogeneous distribution of buoyancy flux in the lake.

Data from a dense array of thermistor chains made possible the

identification of a previously unobserved path of the energy flux in stratified

basins. This path is described in Chapter 4, where the results of the field data

analysis are compared to the results of the mechanistic model that guided their

interpretation.

The data for Chapters 2 and 4 were collected in Lake Kinneret, Israel.

Some of the characteristics of the basin-scale and high-frequency waves in this

lake had been reported over several years of research, but many questions

remained open. The lake has special bathymetric and periodic forcing conditions

that facilitates de study of both basin-scale and high-frequency internal waves.

The data for Chapter 3 was collected in Lake Constance, Germany. Previous

research on basin-scale waves in this lake pointed out the existence of high-

frequency internal waves, opening questions about their role in the energy flux

Page 30: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

6 Chapter 1

path. A sill was believed to play an important role in the generation of high-

frequency internal waves. Both lakes have been focus of extensive research over

the years due to their importance for water supply and the fishing industry.

A summary of the major findings of this work is provided in the

conclusions together with some implications and suggestions for future work.

Page 31: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

7

Chapter 2

Spatial structure of the dominant basin-scale internal waves in

Lake Kinneret

2.1 Abstract

Field data and numerical simulations were used to investigate the effects

of basin shape, continuous stratification, and rotation on the three-dimensional

structure of the dominant natural basin scale internal wave modes in Lake

Kinneret, a stratified lake large enough so that the earth’s rotation influences the

wave motion. The structure of the modes was inferred from power spectral

density of measured and simulated isotherm vertical displacements and from

rotary power spectral density of isopycnal velocities obtained from numerical

simulations. The shape of the lake at the level of the thermocline, in conjunction

with the dispersion relationship, determines the horizontal configuration of the

natural modes. The dominant response to wind forcing was an azimuthal and

vertical mode one Kelvin wave with a natural period of 22.6 h that propagated

around the entire basin; the second most dominant response was a vertical mode

one wave with a natural period close to 10.5 h and composed of two counter-

rotating Poincaré circular cells. The sloping bottom produced an intensification of

vertical displacements and velocities over the slope due to focusing of waves rays

after reflection at bottom slopes close to the critical angle. The amplitude of the

observed oscillations was very sensitive to the phase of the wind relative to the

existing waves; resonance was affected by the cessation time of the wind events

because it defines the local forcing period and resets the phase of the observed

oscillations.

Page 32: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

8 Chapter 2

2.2 Introduction

Basin-scale internal waves are ubiquitous to stratified lakes and reservoirs

and drive a variety of physical, chemical, and biological processes in such water

bodies (Mortimer 1974). They possess most of the energy contained in the

internal wave spectrum and play an important role in energizing both vertical

mixing and horizontal dispersion. Vertical mixing at the metalimnion level erodes

the natural barrier imposed by the stratification and facilitates primary production

in the surface layer by incorporating nutrient-rich hypolimnetic waters (e.g.,

Ostrovsky et al. 1996) into the surface layer. Energy is transferred from the basin-

scale waves to turbulent mixing by sloshing and breaking of basin scale internal

waves around the perimeter of the lake, by the increased shear in the interior of

the lake or by the cascade of energy from the basin scale waves to high frequency

internal waves that are generated either by non linear steepening or by shear

instabilities (Imberger 1998; Boegman et al. 2003). Equally, basin scale internal

waves, combined with the action of the surface wind induced currents, sustain

horizontal dispersion via a host of mechanisms: pseudo-chaotic dispersion

(Stocker and Imberger 2003a), horizontal shear dispersion (Stocker and Imberger

2003b) and dispersion by synoptic eddies formed when internal Kelvin waves

separate at headlands (Ivey and Maxworthy 1992). In addition, the oscillating

motion over a lake bed generates resuspension and transport of sediments,

nutrients, and contaminants (Gloor et al. 1994). An understanding of the temporal

and spatial patterns of the basin-scale internal waves is therefore necessary to

better quantify the above processes that, in turn, determine the success of any

water quality investigation.

Page 33: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 9

Observations of temperature or velocity fluctuations reveal that basin scale

internal waves have characteristic dominant periods, even when generated by

isolated wind events or by wind forcing with different periods (Lemmin 1987;

Saggio and Imberger 1998). The observations also show that the oscillations form

coherent structures along vertical profiles and horizontal transects with length

scales of the order of the dimensions of the basin (Thorpe 1974; Münnich et al.

1992; Hodges et al. 2000). This suggests that lakes are dynamical systems with

natural (or free) modes of oscillations that have characteristic period and spatial

structure and that a knowledge of the characteristics of these modes is crucial for

understanding the baroclinic response of lakes to general forcing.

When the relative importance of Coriolis to buoyancy forces is small, as in

low-latitude or narrow lakes, natural modes are non-rotating seiches and when it

is large, as in large lakes, the natural modes are rotating Kelvin and Poincaré

waves (Mortimer 1974). In either case, the period and the spatial structure of the

natural modes are also functions of the shape and the stratification of the lake.

Analytical solutions for these functions have not been found for general

bathymetries and only approximate solutions exist. Models with simple horizontal

geometries (e.g., rectangular, circular, elliptical) and flat bottoms reproduce the

periods well and give indications of the horizontal structure of the natural modes

for layered (Heaps and Ramsbottom 1966; Csanady 1967; Antenucci and

Imberger 2001a) or for continuous stratifications (Csanady 1972; Turner 1973;

Monismith 1987), but reveal little as to the effect of an irregular basin shape or a

sloping bottom.

More elaborated numerical eigenvalue models allow mode determination

considering a variable depth, but the stratification must be approximated by a

Page 34: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

10 Chapter 2

series of homogeneous layers (Schwab 1977; Salvadè et al. 1988; Bäuerle 1998).

Such numerical schemes yield horizontal structure of the pycnocline

displacements closer to those observed, but the vertical structure of the modes is

poorly reproduced. In particular, the changes of wave number, energy density, and

amplitude of the wave beam on reflection from a sloping bottom (Phillips 1966;

Eriksen 1982; Dauxois and Young 1999) cannot be captured by this type of

models.

The structure of natural modes in a basin with a sloping bottom and a

continuous stratification was studied numerically by Münnich (1996) and Fricker

and Nepf (2000) for non-rotating, vertically two-dimensional basins with variable

depth. They showed that the sloping bottom dramatically changes the spatial

structure of the natural modes compared to those solutions obtained for a

rectangular basin. Fricker and Nepf (2000) pointed out that an extension of their

procedure to 3 dimensions would be difficult.

Simulations with 3D hydrodynamic models can reveal features that can be

attributed to the spatial structure of the natural modes of oscillation. Rueda et al.

(2003) associated features in the horizontal distribution of the integrated potential

and kinetic energy of the vertical mode one, basin scale waves in Lake Tahoe with

major lake boundary irregularities. The difficulty in using 3D modelling to infer

the shape and period of natural modes of oscillation is that, in general, it is

difficult to isolate and interpret their characteristic features when the lake is

constantly forced by an unsteady surface wind stress.

In Lake Kinneret (Israel), 24 and 12-h-period signals dominate the power

spectra of the observed isotherm displacement and horizontal velocity (Serruya

1975; Ou and Bennett 1979; Antenucci et al. 2000). Serruya (1975) attributed the

Page 35: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 11

existence of these periodic motions to the interaction of the periodic wind with a

12-h natural period of oscillation, and documented the cyclonic rotation of the 24-

h-period component of the signal. Ou and Bennett (1979) described the 24-h-

period oscillation as a vertical and horizontal mode one (V1H1) Kelvin wave, and

suggested that the 12-h-period motion was a non-rotating V1H1 seiche with a

nodal line in the north-south direction. With the circular flat bottom basin model

approximation (Csanady 1967), Antenucci et al. (2000) estimated the period of

the natural internal modes in Lake Kinneret to be 22.5 and 12.2 hours and

interpreted these as V1H1 cyclonic Kelvin and anticyclonic Poincaré waves

respectively. The proximity to the periods observed in the field led them to

suggest that these two waves were the dominant modes in the lake. In a later paper

Antenucci and Imberger (2003) showed how these modes moved in and out of

resonance with the 24-h wind forcing as the stratification changed over a seasonal

time scale (I refer to resonance as the amplification of the motion that occurs

when the period and spatial structure of the forcing coincide with those of one of

the natural modes of the system, and not in the sense used by Maas and Lam

(1995) who called a resonance the condition in which all wave rays close upon

themselves after multiple reflections at the boundaries of the domain). The

numerical simulations of Hodges et al. (2000) confirmed the rotational character

of the 24-h component, but they attributed the 12-h component to a seiche across

the EW axis of the lake with a small NS component. The conflicting results from

these studies indicate that the 12-h component remains to be unequivocally

characterized. Also, the dependence of the modal structure on the bathymetric

shape, which is of critical importance in determining the bottom and metalimnetic

shear stress distribution, remains to be determined.

Page 36: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

12 Chapter 2

To address these issues, a field campaign was undertaken during a typical

summer stratification period. After describing the field campaign, we illustrate the

effects of the horizontal shape and sloping bottom of the basin on the structure of

the dominant natural modes by using spectral analysis of field data and results of

complementary numerical simulations. We also estimate their natural periods and

show that the wind resets the phase of the observed wave field, imposing the

observed periods, and that the time of termination of the wind events is a very

important parameter for the observed resonant interaction with the wave field.

2.3 Field site

Lake Kinneret is a warm, monomictic lake with the turnover occurring in

late December or early January. The summer stratification is characterized by a

thermocline 16-18 m deep and a temperature difference between the epilimnion

and the hypolimnion of up to 8°C (Serruya 1975). The lake is approximately 22

km long north-south and 15 km wide west-east and the maximum depth was about

39 m during the summer of 2001, when the field experiment for this research was

conducted (see Fig. 2.1). During the stratified condition, when internal wave

activity is strongest, the dominant component of the wind forcing is a strong

westerly breeze that blows every afternoon (Serruya 1975). Inflows, outflows, and

density variation due to salinity have a negligible effect on the dynamics at this

time of the year.

2.4 Field equipment

To record the internal waves activity, 10 Lake Diagnostic Systems (LDS),

containing thermistor chains and wind velocity sensors, were deployed in Lake

Page 37: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 13

Kinneret during the summer of 2001. Five of the LDS’s were located close

together to identify the characteristics of short wavelength internal waves and the

results from these chains will be described elsewhere; only data from one of these

stations (Tv) is considered here. The location of the 6 LDS stations used for the

present analysis is shown in Fig. 2.1 and details of their deployment are given in

Table 2.1. The accuracy of the thermistors was 0.01°C (with resolution of

0.001°C). Stations Ty (one of the stations close to Tv) and T9 also had full

meteorological stations mounted 1.5 m above the water surface.

2.5 Observed wind pattern

Each day during the experiment, westerly winds started abruptly after

midday, reached speeds up to 11 m s-1 that were maintained for 6 to 8 hours (Fig.

2.2 a,b), similar to that observed previously by Serruya (1975). Power spectral

density (PSD; Bendat and Piersol 1986) of the wind speed (Fig. 2.2c) reveals that,

in addition to the dominant 24-h periodicity, the wind signal contained a

secondary 12 h periodic component that had also been identified by Ou and

Bennett (1979) and Antenucci and Imberger (2003) in previous field campaigns.

The spatial variability of the wind field during the 2001 field experiment was

documented by Laval et al. (2003b); the effects of the daily variations in the wind

pattern are discussed below.

2.6 Measured basin scale internal waves activity

The isotherm displacements (Fig. 2.3) and the PSD of the vertical

displacement of the 23°C isotherm (Fig.2.4), located in the middle of the

metalimnion, show that a 24-h-period oscillation was dominant at all stations.

Page 38: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

14 Chapter 2

This oscillation had larger amplitude at the stations located along the 20 m depth

contour (stations Tg, Tf, T7, and T9) and progressively smaller amplitude towards

the interior of the lake (stations Tv and T4). The second largest spectral peak

corresponds to a 12-h-period oscillation, which was larger at stations Tf, T9, and

Tv and just significant at stations Tg, T7, and T4. The mean stratification in which

these waves propagated was obtained by averaging the elevation of the isotherms

at station T4 over 24 h from day 170.5 to 171.5 and is shown in Fig. 2.5 together

with the corresponding buoyancy frequency profile. The buoyancy frequency was

everywhere, even in the relatively well mixed epilimnion and hypolimnion, three

orders of magnitude larger than the frequencies of the two dominant oscillatory

motions (1×10-5 and 2×10-5 Hz). Therefore, the thermocline did not form a wave

guide with turning levels and the corresponding wave rays could propagate in the

entire lake from the surface to the bottom. The horizontal variation of the time

averaged temperature profile and any associated consequences could not be

investigated because the metalimnion intercepts the bottom of the lake at the

location of the stations Tg, Tf, T9, and T7.

2.6.1 Propagation

The propagation of the 24 and 12-h-period waves along the perimeter of

the basin was tracked by following their phase at the stations located along the 20

m depth contour. To do so, the displacement of the 23°C isotherm at each of the

stations (Fig. 2.6a) was band-pass filtered around 24 h (between 19.8 and 27.8

hours; Fig.2.6b) and around 12 h (between 10.7 and 13.9 hours; Fig.2.6c) using a

fourth order Butterworth filter. The crests of the 24-h-period signal is seen to

Page 39: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 15

propagate cyclonically (anti-clockwise) with oscillations at Tg and T7, located at

opposite ends of the basin, out of phase, a characteristic of a basin-long azimuthal

mode 1 wave. By contrast the 12-h oscillation propagated anticyclonically

(clockwise; Fig. 2.6c), but the oscillations were almost in phase at stations Tg and

T7. This suggests that this oscillation did not correspond to a mono-cellular

anticyclonic wave of the first azimuthal mode filling the entire basin, as

postulated by Antenucci et al. (2000). It is possible that wave crests propagated

along the transect T9-Tf-Tg while crests at T7 followed a disconnected path. A

different line type is used in Fig. 2.6c to indicate that the propagation of wave

crests between T7 and T9 was questionable.

2.6.2 Azimuthal structure

We studied the azimuthal structure of the dominant components of the

wave field by looking at the evolution, over a period of 24 h, of the band-passed

signals of the 23°C isotherm at the stations located along the 20-m depth contour.

The evolution of the 24-h component (Fig. 2.7a) resembles the cyclonic

propagation of an azimuthal mode 1 wave covering the entire basin. The 12-h

component (Fig. 2.7c) appeared to be of an azimuthal mode 2 wave propagating

anticyclonically around the basin with oscillations at stations Tg and T7 in phase.

However, data recorded at 5 stations along a West-East transect in 1999 (not

shown) suggest that the 12-h component was azimuthal mode one, as does the

transect T9-Tf-Tg in Fig. 2.7c. This leads to the hypothesis that the structure of

the 12-h oscillation was dominated by an azimuthal mode 1 cell in the main part

Page 40: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

16 Chapter 2

of the basin which did not reach the southern embayment of the basin where

station T7 was located.

2.6.3 Radial structure

The phase along a radial transect defined by stations Tf, Tv, and T4 and

the monotonic decrease of amplitude with distance from the shore (Fig. 2.7b, d)

indicates that the two dominant components of the wave field were of the first

radial mode.

2.6.4 Vertical structure

The vertical modal structure of the 24 and 12-h-period oscillations was

identified by calculating the phase of the vertical displacement over the depth of

the thermistor chains. The 23°C isotherm was chosen as reference for the phase

and the areas where the isotherms intersected the bottom, at any time, were

excluded from the analysis. Fig.2.8 shows the results at those stations where the

amplitude of both the 24 and 12-h oscillations were large enough to make the

coherence and phase statistically significant. The small change in phase over the

water column indicates that a vertical mode one dominated both periodic

components.

The cyclonic propagation of phase of the 24-h motion and its horizontal

and vertical spatial structure suggests that the motion was a Kelvin wave with a

vertical and horizontal mode one structure that extended over the whole basin,

analogous to that for a circular (Antenucci et al. 2000) or elliptical (Antenucci and

Imberger 2001a) basin shape; the bathymetry did not influence the modal

Page 41: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 17

structure of this wave. In contrast, the bathymetry appeared to have a dramatic

influence on the modal structure of the 12-h wave, the field data suggesting a

vertical mode one wave with a horizontal structure with two cells. The first cell, in

the main part of the basin, seems to have a radial and azimuthal mode one

structure with anticyclonic propagation of phase; the structure of the second cell,

located in the embayment in the south, could not be resolved with the field data,

and was examined using three-dimensional numerical simulations.

2.7 Numerical simulations

We carried out numerical simulations with the Estuary and Lake Computer

Model (ELCOM) to increase the spatial resolution of the temperature information

and to derive the velocity field. Only a brief description of the model is given here

and the reader is referred, for further details, to the original publications by

Hodges et al. (2000). The model solves the 3D, hydrostatic, Boussinesq,

Reynolds-averaged Navier-Stokes equations and scalar transport equations of

potential temperature, salinity and tracers in a Z-coordinate system. Diffusion and

advection are separated both for momentum and scalars and a mixing model,

based on the integral solution of the turbulent kinetic energy equation, is used to

estimate the vertical diffusive transport (Simanjuntak, M. A. pers. comm.). The

model also includes a filter to control the numerical diffusion of potential energy

(Laval et al. 2003a), so the stratification and the phase speed of the internal waves

are not altered numerically over the simulation period. Heat and momentum

exchange at the water surface was estimated by bulk transfer models (e.g.,

Amorocho and DeVries 1980; Imberger and Patterson 1981; Jacquet 1983), the

transfer coefficient being corrected for the stability of the atmospheric boundary

Page 42: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

18 Chapter 2

layer (Hicks 1975). The surface energy fluxes are separated into a non-penetrating

component introduced in the surface mixed layer, and a penetrating component

introduced over one or more layers according to Beer’s law. Validations with

temperature (Hodges et al. 2000; Laval et al. 2003a) and with velocity data

(Marti, C pers. comm.) showed that ELCOM reproduces well the internal wave

dynamics in Lake Kinneret. Before using the model to investigate the effect of the

bathymetry on the structure of the natural modes, we tested its ability to reproduce

the propagation and vertical structure of the dominant observed oscillations for

the 2001 data set.

2.7.1 Validation

The bathymetry of the lake was discretized with a 400×400×1 m grid, and

a time step of 450 s was used to satisfy the CFL stability condition for the internal

motions. Short wave radiation, net radiation, air temperature and relative humidity

inputs were averaged every 10 minutes from direct measurements taken every 10

seconds at the meteorological station Ty. A 2D wind field was obtained through

spatial linear interpolation of the records from the stations in and around the lake

(see Fig. 2.1). More details about the interpolation of the spatially variable wind

may be found in Laval et al. (2003b). The initial temperature profile is taken from

Fig. 2.5a and is considered horizontally homogeneous over the lake. The spin-up

time of the model was 2 days.

The spectra in Fig.2.4 demonstrate that the model reproduced the

concentration of energy at periods of 24 and 12 h. Figure2.9 shows the time series

of the 23°C isotherm displacement and its band-pass filtered components at the

six stations. It reveals that the model underestimated, by about 20%, the amplitude

Page 43: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 19

of the 24-h component, but simulated well its phase, except at station T4 where

the small amplitude of the oscillation made the coherence and phase analysis

statistically not significant. We show below that the underestimation of the

amplitude was due to artificial dissipation introduced by the no normal flow

boundary condition applied on the step-like bottom discretization, and that this

weakness of the model does not affect the phase and the modal structure of the

simulated waves. The phase of the 12-h component was also well simulated at the

stations away from the north-south axis (Tf, Tv, and T9) where the coherence and

phase analysis were statistically significant. The ability of the model to reproduce

their propagation around the basin is verified through an analysis of coherence

and phase. The results are summarized in Table 2.2 and show that both periodic

signals were coherent in the stations around the basin in the field measurements

and also in the model results. The model results were also coherent with the field

data at most of the stations. The cyclonic propagation of phase of the 24-h signal

from Tg to T7 and anticyclonic propagation of phase for the 12-h signal from T9

to Tg were reproduced well by the model, with relatively small differences of both

components between the model and the field data. Model results were almost out

of phase with field data at T4 but these results were statistically not significant.

Excluding T4, the station where the biggest relative phase difference occurred for

both signals was T7, the difference being 1.7 hours (7% of the observed period)

for the 24-h signal and 2 h (17%) for the 12-h signal.

The vertical structure of the model results was analysed by calculating the

phase of the isotherm displacements over the depth of the simulation grid at

locations equivalent to those of the thermistor chains. The results (Fig. 2.8)

reproduced well the dominant vertical mode 1 structure of both periodic

Page 44: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

20 Chapter 2

components observed in the field data. Higher vertical modes might still be

present in both the field data and the simulation results, but due to the smaller

potential to kinetic energy ratio associated with these modes (Antenucci and

Imberger 2001a) they did not contribute significantly to the isotherm

displacement.

2.7.2 Numerical damping in ELCOM

To show that the extra damping in the ELCOM was due to the step

bathymetry, we estimated the decay rate of the oscillations in the field and in the

model and calculated the relative amplitude of resonant oscillations in two

oscillating systems with the different decay rates. The decay rate in ELCOM was

estimated by visually fitting a decaying exponential envelope curve to oscillations

that follow the release of a tilted stratification in runs with different grid sizes.

The associated e-folding times are compared in Table 2.3. The same procedure

could not be applied to the field data as the lake was forced continuously by the

wind before any decay could be evaluated. Instead, we determined the decay rate

and the natural period of the Kelvin wave (the most energetic wave) by matching

the 24-h band-pass filtered oscillation of the 23°C isotherm at Tv to the oscillation

of a linear damped forced mass-spring system. This simple dynamic system

describes the oscillations at a particular location of a lake without the complexity

added by the spatial structure (Stocker and Imberger 2003a). The forcing, F, is

given by the input of wind momentum approximated by 30-minute long uniform

wind blocks. In turn, each constant wind block was considered as made up of a

suddenly imposed steady wind event of infinite duration and an equal but negative

wind event starting 30 minutes later. At a particular location in the basin, the

Page 45: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 21

isotherm oscillation generated by every wind event (positive or negative) is given

by:

{ } ( )lag(t t )i i lagx A F 1 cos (t t ) e H t−α −⎡ ⎤= − ω −⎣ ⎦ (2.1)

where

2 20

c ,2m

2α = − ω = ω − α , (2.2)

x is the displacement from the equilibrium position, t is the elapsed time from the

start of the wind, c is the effective damping coefficient, the coefficient A and the

phase account for the spatial location and H is the unit step function. The

parameter is the inverse of the e-folding time,

lagt

α ω is the damped natural

frequency and is the undamped natural frequency of the system. The

displacement generated by the wind series is given by linear superposition:

, (2.3) ( ) (M

lag i sii 1

X t, , ,A, t x t t=

α ω = −∑ )

where is the time of start of the i-th wind event and M is the total number of

positive and negative wind events. The values of

sit

α and ω were obtained by

least-squared fitting X with the 24-h band pass filtered 23°C oscillation at station

Tv. The 3.9-day decay time scale for the amplitude of the Kelvin wave, α ,

translates into an energy decay time scale of 2 d, which is about half that

estimated by Antenucci and Imberger (2001a). This value is also well below the

11.6 d estimated from the experimental results of Wake et al. (2005) in a 2-layer

flat bottom circular basin with the Burger number and inertial period similar to

that found in Lake Kinneret, where the transference of energy to smaller scales is

limited by the absence of topographic features. From the natural frequency, , a ω

Page 46: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

22 Chapter 2

natural period of 22.6 hours (see Table 2.3) was calculated for the free Kelvin

wave.

Using Eq. 2.2, we estimated 0ω , leading to an associated natural period in

the absence of damping, , of 22.59 h. With 0T 0ω and the different decaying rates,

we estimated for several grid sizes; the associated periods (Table 2.3) reveal

that the effect of the extra damping on the damped natural period was minor.

Although the propagation speed of the simulated Kelvin wave depends on the grid

size (Bennett 1977; Beletsky et al. 1997), the effect is practically eliminated if, as

in all our simulations, the grid size is less than about one-fifth of the internal

radius of deformation (Schwab and Beletsky 1998).

ω

The resonant amplitude of the long-term oscillation excited by a harmonic

force of amplitude and frequencyFA Fω , given by:

( )0.521 2 2 2 2

F 0 F FB A m 4−

− ⎡ ⎤= ω − ω + α ω⎢ ⎥⎣ ⎦, (2.4)

(Table 2.3) shows a 23% reduction of amplitude for the 400×400 grid when the

forcing period is 24 h. This is representative of what is observed in Fig. 2.9,

indicating that the smaller amplitude in the model is an effect of grid step

damping.

From the above we conclude that, despite numerical damping, the model

provides a valuable tool to expand the characterization of the modal shape of the

isotherm displacements that are induced by the dominant basin scale internal

waves.

Page 47: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 23

2.7.3 Basin with flat bottom and irregular horizontal shape

We took the analytical solution for the Kelvin and Poincaré natural modes

in a circular flat-bottom basin (Csanady 1967, 1972; Antenucci et al. 2000) as the

starting point in our investigation of the effects of the irregular bathymetry on the

shape of the natural modes. For elliptical basins, Antenucci and Imberger (2001a)

demonstrated that azimuthal mode one Kelvin waves have maximum isotherm

displacements where the radius of curvature of the basin is minimum and larger

velocities where the radius of curvature is smaller. This behaviour is reversed for

Poincaré waves.

To look at the effect of the more irregular basin perimeter, we carried out a

simulation in an idealized 30 m deep basin with flat bottom, the horizontal shape

of the 15 m depth (mean depth of the thermocline) contour in Lake Kinneret at

(Fig. 2.1) and the stratification of the top 30 m of the time-averaged temperature

profile in Fig. 2.5a. A 400×400×1 m grid and a 450 s time step were used. The

thermodynamic module in ELCOM was switched off and no wind forcing was

applied, allowing the system to oscillate freely from an initial west-east tilt of the

stratification. Temperature and horizontal velocity time series were recorded at

every grid point and used to calculate isotherm displacements and isopycnal

velocities time series. The power spectral density (PSD) and rotary power spectral

density (RPSD; Gonella 1972) of these time series were then calculated and used

to identify spatial variations in the amplitude of the oscillations. The PSD and

RPSD revealed dominant periodicities close to 23 and 10.5 hours for the vertical

displacements and isopycnal velocity. Figure 2.10 shows the energies of the PSD

of vertical displacement and of the RPSD of isopycnal velocities for isotherms in

the epilimnion (26.5°C), metalimnion (23°C), and hypolimnion (17°C). The 23-h

Page 48: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

24 Chapter 2

signal has the characteristics of an azimuthal and vertical mode one wave: larger

displacements toward the shore (Fig. 2.10a), especially in the metalimnion, and

larger cyclonic velocity component in the epilimnion and hypolimnion with

almost no motion in the metalimnion (Fig. 2.10b). The elongated shape of the lake

produced an increase of the vertical displacement where the radius of curvature

was minimum and an increase of the velocities where the radius of curvature was

maximum, similar to that found for an elliptical basin by Antenucci and Imberger

(2001a). The vertical displacement of the 10.5 h period wave was larger towards

the perimeter, specially at three locations in the northwest, east and southwest

(Fig. 2.10c), while two amphidroms were located in the middle of the basin and in

the southern embayment. Figure 2.10d shows the difference between the

anticyclonic and cyclonic components of the velocity at 10.5 hours. Positive

values indicate anticyclonic (clockwise) rotation; negative values, cyclonic

(anticlockwise) rotation. Two vertical mode one counter-rotating cells combined

to make the horizontal structure. The first cell filled the main body of the basin

and had a classic Poincaré type structure with velocity vectors rotating

anticyclonically and the largest magnitude in the centre. In the second cell, located

in the embayment in the south of the basin, the velocity vectors rotated

cyclonically. The location of Station T7 in the southern cell (Figs. 2.1, 2.10d) thus

explains why we did not observe a simple structure in the field data when

investigating the propagation of phase around the entire basin.

2.7.4 Lake bathymetry with variable depth

To include the effect of the sloping bottom on the spatial structure of the

natural modes of oscillation, we simulated the motion following an initial tilt of

Page 49: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 25

the stratification with the real bathymetry of the lake; all other simulation

variables were kept the same. The energies of the PSD and RPSD at the dominant

periods of 23 and 10.5 hours are presented in Fig. 2.11. All those areas where the

isotherms intersected the bottom at any time were excluded from the PSD and

RPSD plots so details of the structure close to the edges of the lake were lost.

Despite this, some features still provide an insight of the effects of the sloping

bottom on the structure of the natural modes. The 23-h-period wave had larger

energy (PSD) of the 23°C isotherm displacements towards the north and south

edges of the basin (Fig. 2.11a) similar to the flat bottom, irregular horizontal

shape case. In the hypolimnion, in contrast to the small displacements observed

for the flat bottom case, large vertical displacements occurred over the mild

bottom slopes in the south and the west of the lake, revealing an intensification of

the motion over the sloping bottom. In the hypolimnion (17°C isotherm), the

cyclonic component of the RPSD of velocity at 23-h (Fig. 2.11b) was also larger

toward the south and west. Intensification of the isotherm vertical displacement

and isopycnal velocity over the sloping bottom also occurred for the 10.5-h-period

wave (Fig. 2.11c,d). The circular Poincaré type structure of flat bottom basins

with larger velocities in the middle of the cell was retained in the epilimnion

(26.5°C isotherm). The two-cell structure of this mode was difficult to observe for

the sloping bottom case because the resolution of the method, as mentioned

above, was not sufficient close to the boundary. However, a vertical cross-section

along the north-south axis of the lake (Fig. 2.12) shows the anticyclonic rotation

of the velocity in the main part of the basin and the cyclonic rotation in the south

end. The vertical mode one was maintained, but the surface that separates the two

cells was tilted.

Page 50: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

26 Chapter 2

2.8 Interaction of the wind and the dominant natural modes

The field observations reveal that the observed dominant periods in the

isotherm displacement are the dominant periods in the forcing and not the period

of the natural modes (Figs.2.2c and 2.4). We suggest that the observed oscillations

result from the superposition of the free internal waves that are generated by

successive wind events. Here, we illustrate how superposition of natural modes

explains the effects of the daily changing wind pattern on the observed

oscillations.

2.8.1 Reset of the phase and modification of the period

With each onset of the wind (Fig. 2.13a), epilimnetic water started to

move to the southeast, the wind forcing it eastward and the Ekman transport

turning the flow southward, creating upwelling in the Northwest between stations

Tf and Tg (see also Ou and Bennett, 1979). At Tf, downwelling of the 23°C

isotherm changed into upwelling with the onset of the wind on days 170 and 177;

on days 171 and 173 to 176, the isotherm was already rising at the time of the

onset of the wind, but the rate of upwelling increased when the wind started (Fig.

2.13b). While the wind was pushing the surface water eastward, a thermocline

slope across the basin was built with associated upwelling at Tf and downwelling

at T7 (Fig. 2.13b,c), adding to the one set up by the existing wave field.

Regardless of the phase of the preceding wave field, a new set of waves was

released at the end of every wind event: strong downwelling at Tf and upwelling

at T7 began when the wind fell below 5 m s-1 (Fig. 2.13b,c). The new set of waves

released at the end of every wind event added to the waves already propagating

around the lake and the combination appears as a reset of the phase of the

Page 51: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 27

observed wave field and overrides the natural period of 22.6 h with the mean

forcing period of 24 h.

2.8.2 Resonance

The effect of the temporal pattern of the wind on the wind-wave field

interaction was investigated using extended records between days 169 – 205 (18

June – 24 July) at stations Tg, Tf, and T7. We concentrated on the most energetic

24-h signal. The parameters determining the amplitude of a forced oscillating

system are the magnitude of the forcing, the forcing to natural periods ratio and

the actual to critical damping coefficients ratio (Norton 1989). In general, lakes

are under-damped oscillating systems (Mortimer 1974), for which the maximum

resonant amplitude is reached when the forcing to natural periods ratio approaches

1 (Norton 1989). The magnitude of the forcing is given by the input of momentum

from the wind:

e

s

t

t

1Im p dt= τρ ∫ (2.5)

where

D airC Uτ = ρ U (2.6)

is the shear stress over the water surface, and are start and end times of the

wind event, is the water density, t is time, is a drag coefficient, is the air

density and U is the east-component of the wind velocity at Tf. As a new set of

waves is released at the end of each wind event, we considered the time from the

end of the previous wind event to the end of the wind event under examination,

Δt, as the local period of the forcing. The daily variability in the characteristics of

st et

ρ DC airρ

Page 52: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

28 Chapter 2

the wind forcing and amplitude and time of the peak of the 24-h signal at Tg, Tf,

and T7 is presented in Fig. 2.14.

At Tf, the amplitude was about 3 m (Fig. 2.14d) and the crests reached the

station at around 20 h (Fig. 2.14e) for the period prior to day 190; after that date,

the amplitude dropped to values around 1.2 m due to a decrease in the input of

wind momentum (Fig. 2.14a), and the crest reached Tf at about 19 h as a direct

result of the earlier termination of the wind events (Fig. 2.14b). On days 176, 187,

and 188, Δt was close to the natural period of the Kelvin wave (Fig. 2.14c), so the

wave being generated was in phase with the wave generated by the previous wind

event and hence the amplitude of the internal waves increased (Fig. 2.14d). The

strong winds on days 174, 175, 185, and 190 did not produce large amplitude of

the internal wave because Δt was far from that required for resonance.

The gradual positive phase shift of the wave crests (Fig. 2.14e) was the

effect of the natural period of the wave being shorter than the average period of

the wind and was accentuated when the momentum introduced by the wind was

small as was the case on days 171 to 173. Events of strong wind momentum input

generated a negative phase shift (days 174 to 176, 185 to 186), especially when

the end of the wind event was delayed as on days 175 and 186.

2.8.3 Spatial structure of the observed oscillations

The spatial structure of the internal wave oscillations observed under

atmospheric forcing was extracted from the simulation used for the validation.

The complete structure cannot be visualized because of the lack of resolution of

our method close to the bottom, but the results permit the interpretation of the

PSD and RPSD as a combination of the signatures of the vertical mode 1 natural

Page 53: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 29

modes described above with other motions excited by the wind: The 24-h-period

signal (Fig. 15a,b) shows elements of the free Kelvin wave (Fig. 2.11a,b) such as

the larger displacements toward the north and south ends of the lake and the

intensification of motion over the sloping bottom in the west. The strong vertical

displacements observed in the epilimnion at the west and east ends on the lake is

the signature of upwelling and downwelling at the ends of a surface layer

circulation sustained while the wind was blowing. The 12-h-period signal (Fig.

2.15c,d) showed strong currents and vertical displacements over the sloping

bottom as in the 10.5-h-period natural mode (Fig. 2.11c,d). However, additional

features in the PSD and RPSD structures suggest that other natural modes with

periods close to 12 h were excited. The most noticeable one, a patch of high

anticyclonic rotating velocities observed in the metalimnion (23°C isotherm) in

the north of the basin suggests that the wind also excited a vertical mode 2 wave

(Lighthill 1978). The existence of this mode is supported by the rotary spectra of

velocity data collected close to the location of Tv in 1998 (Antenucci et al. 2000).

The intensification of the velocity over the sloping bottom has, so far, been

identified only from model results. We now show that the same slope interaction

is seen in the field data by calculating the profile of RMS vertical displacements,

, at station Tv (Fig. 2.16), located on the western slope. To do so, we

calculated (following Fricker and Nepf 2000), at the elevation of each thermistor,

the time series of temperature fluctuation with respect to the mean temperature

profile in Figure 2.5 and band passed this at the dominant observed periods, 24

and 12 h. Assuming that changes in temperature are due mainly to vertical

displacement (i.e., assuming very small horizontal temperature gradients) the

vertical displacement is give by:

rmsς

Page 54: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

30 Chapter 2

j

j rmsrms

TT z

Δς =

∂ ∂ (2.7)

where is the RMS temperature fluctuation for the 24-h (j = 1) and 12-h (j =

2) waves and

jrmsTΔ

T z∂ ∂ is the mean temperature vertical gradient. The surface mixed

layer is excluded because the temperature fluctuations there are dominated by the

heating/cooling diurnal cycle rather than by wave induced vertical motions.

Clearly seen from Fig. 2.16 is the amplification of the vertical motion over the

gently sloping western slope, which is reproduced by ELCOM reasonably well.

2.9 Discussion

Temperature data showed that, unlike the Kelvin wave, the 12-h

component of the observed oscillations did not have the structure of a wave

propagating around the entire basin, pointing to an effect of the irregular

bathymetry. Two vertical mode one counter-rotating cells developed as the result

of the plan shape of the basin. This conclusive result embraces previous evidence

about the identity of the 12-h signal (Ou and Bennet 1979; Hodges et al. 2000;

Antenucci et al. 2000).

There is previous numerical evidence of internal Poincaré waves being

locked to some features of the basin (e.g., Wang et al. 2000) and of internal

natural modes with a multi-cellular horizontal structure (Schwab 1977; Lemmin et

al. 2005); we offer here a dynamical explanation for their occurrence in terms of

the dispersion relationships for circular basins developed by Antenucci and

Imberger (2001a). The structure of the 21°C isopycnal velocity of the 10.5-h-

period wave (Fig. 2.17) shows circular cells with radii, r0, of 4800 and 3000 m

separated by a no-flow across vertical cross-section. A layered model (Csanady

Page 55: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 31

1982; Monismith 1985) with layer temperatures of 26.5, 21.5, and 16.5°C and

depths of 12.5, 11.0, and 6.5 m gave a first vertical mode phase speed (c1) of 0.32

m s-1. With f, the local inertia frequency, equal to 7.81×10-5 rad s-1, the first

vertical mode Burger numbers (S1 = c1 r 0-1 f-1) of the large and small cells were

0.85 and 1.36 respectively. The ratio of the inertial period (22.3 h) to the natural

period of this mode (10.5 h) was 2.1. Figure 2.18a shows that the waves in both

cells satisfied the dispersion relationships for azimuthal mode 1 counter rotating

Poincaré waves in circular basins. This results suggest that the horizontal structure

of a natural mode has several rotating cells if circular (or elliptic) sub-basins

holding rotating waves with the same period, vertical and azimuthal modes can

geometrically fit into the shape of the basin at the level of the thermocline. Every

cell holds a rotating wave as it were in an individual basin; the size of the basins

being such that the periods of the waves in all the cells are the same. The ratio of

the radii of the anticyclonic and cyclonic cells coexisting in the horizontal

structure of a natural mode, calculated from the dispersion relationships, is thus a

useful parameter to predict the existence of counter-rotating cells in a natural

mode. The ratio for azimuthal mode 1 cells is presented in Fig. 2.18b as a function

of the Burger number of the anticyclonic cell, which is always larger and should

occupy the main area of a lake.

Both vertical one natural modes showed a magnification of vertical

displacements and velocities over the sloping bottom, similarly to what Münnich

(1996) and Fricker and Nepf (2000) found for non-rotating seiches in two-

dimensional basins with variable depth. Fricker and Nepf (2000) showed that the

magnification of currents and the location of a definite amplitude maximum

somewhere over the middle of the slope depends on the structure of the

Page 56: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

32 Chapter 2

stratification, unlike the flat bottom case where the maximum of bed velocity is

always located in the middle of the basin. In similar fashion, the maximum bed

velocity in the 3D case is located away from the middle of the cell, where it would

occur in a flat-bottom Poincaré wave (Csanady 1967; Antenucci et al. 2000). The

amplification of the motion over the sloping bottom suggests that the common use

of eigenfunctions calculated under the assumption of a flat bottom and negligible

vertical bottom velocity (e.g., Monismith 1987; Münnich et al. 1992; Boehrer

2000) should be applied with caution when estimating the vertical structure of the

natural modes in lakes.

The intensification of the motion over the sloping bottom may be the result

of focusing of internal waves rays after reflection from the lake bed, which is

larger if the bottom slope is close to the wave ray slope (the critical slope; Phillips

1966; Dauxois and Young 1999). Despite the step-like structure representation of

the bathymetry in ELCOM, the effects of focusing of the wave rays can be

simulated, but limited to scales larger than the grid size; the magnitude of the

local focusing is better simulated when the slope between bottom cells centres is

similar to the real slope.

The slope of the rays of the 10.5 h period wave at the bottom, sgb, may be

estimated from the dispersion relationship for free propagating superinertial

internal waves under rotation (LeBlond and Mysak 1978) as

2 2

2gb 2

b

sNω −

= 2− ωf (2.8)

where is the frequency of the natural mode and Nω b is the buoyancy frequency at

the bottom of the water column. The interaction with the sloping boundary

generates refraction of these waves but the slope of the rays remains unchanged

Page 57: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 33

(LeBlond and Mysak 1978). The ray slope of the subinertial internal Kelvin wave

cannot be described by Eq. 2.8. Rhines (1970) showed that, as for internal Kelvin

waves propagating along a vertical and straight wall, the dispersion relationship of

subinertial internal waves propagating along a wall tilted at any angle from the

vertical is that of free internal waves in the absence of rotation. The inertial

frequency, f, only influences the trapping of the motion to the wall. Based on

these results, we estimated the slope of the rays of the Kelvin wave at the bottom

from the dispersion relationship of free waves in the absence of rotation as

2

2gb 2

b

sN

ω= 2− ω

(2.9)

As seen from Figs. 2.19 and 2.11c,d, the ray slope was similar to the bottom slope

close to the areas where intensification of the oscillations was observed in the

west and, for the case of the 22.6-h Kelvin wave, in the south. Therefore, it is

likely that the observed intensification over the slope was the signature of the

propagation of wave rays that were strongly focused after reflection in the bottom.

We based our analysis of the observed oscillations in terms of resonant

interaction of the wind with the natural modes of the basin. Using geometrical

tracing of wave rays, Maas and Lam (1995) showed that, in an inviscid 2D

parabolic basin with constant buoyancy frequency wave rays do not close upon

themselves and natural modes do not exist. Instead, for some frequencies (and

their associated angles of propagation), focusing makes all the wave rays

converge toward a fixed trajectory called an attractor where all the energy

concentrates and coherent modes are no longer observed. For other frequencies,

the focusing is balanced by defocusing and the rays do not converge toward an

attractor but do not close either, producing what they defined as an infinite period

Page 58: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

34 Chapter 2

attractor where one ray fills the entire basin. These authors suggest that natural

coherent seiching modes in sloping bottom basins are possible only for particular

geometries.

Their analysis relies on the geometric construction of the characteristics

using the dispersion relationship for linear inviscid internal waves in a medium

with constant buoyancy frequency, N, where the angle that the wave ray forms to

the horizontal, , is maintained after reflection from a slope (Phillips 1966). If

viscosity is considered, the resulting dispersion relationship is

θ

(2.10) 2 2 2 2i K N cos 0ω + ων − θ =

where is the kinematic viscosity and K is the magnitude of the wave vector

(Kistovich and Chashechkin 1995). After reflection, changes in K must then be

compensated with changes in θ , especially when focusing of the wave rays

produces larger values of K. For low-period attractors, viscous diffusion then

balances geometrical focusing of the rays and the motion concentrates in a shear

layer around the attracting trajectory, as is shown numerically by Dintrans et al.

(1999) and Rieutord et al. (2001). These authors also showed that regular coherent

modes are recovered from the infinite-period attractors when viscous diffusion is

taken into account.

ν

Münnich (1996) found numerically a V1H1 natural mode in a parabolic

basin for which Maas and Lam (1995) ruled out the existence of any natural

mode. Although viscous diffusion was not considered explicitly in Münnich’s

work, numerical diffusion may have had a similar effect to its physical

counterpart, leading to the occurrence of natural modes. Besides viscous

diffusion, nonlinear effects (Dauxois and Young 1999; Manders and Maas 2004)

and the breakdown of the hydrostatic approximation (Maas 2001) can also

Page 59: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 35

invalidate the conclusions in Maas and Lam (1995) about the non existence of

coherent natural seiching modes. Our data and previous observations (e.g., Thorpe

1974; Lemmin 1987; Münnich et al. 1992) suggests that natural modes of

oscillations are ubiquitous in stratified lakes and focusing and defocusing of all

internal wave rays must balance to produce wave rays that close upon themselves.

Maas et al. (1997), Maas (2001) and Manders and Maas (2004)

demonstrated in the laboratory the existence of shear layers around the predicted

trajectory of a low-period attractor. We believe that they are not observed in

stratified lakes for several reasons. First, the wind forcing is not a pure harmonic

motion (or a simple combination of few pure harmonic motions) sustained for a

high number of wave periods, as was always the case in the mentioned

experiments. Second, the long forcing time necessary to generate a low-period

attractor also indicates that they are highly dissipative and that it is very unlikely

that they remain after the forcing ceases. This was observed by Boegman (pers.

comm.), who noticed that forced interfacial modes generated by harmonically

forcing a two-layer rectangular tank die much faster than a period of the gravest

mode after the forcing stops. In addition, the long scale of the wind forcing at the

lake surface creates over-specification of the boundary conditions (Maas and Lam

1995) and destroys the attractor. Manders and Maas (2003) reported difficulties to

observe attractors when a length scale is imposed by an oscillating paddle.

We described the observed oscillations in Lake Kinneret as the

superposition of internal natural modes generated by every wind event, on

average, every 24 hours rather than as a harmonic forced internal wave with a 24-

hour period. This representation is more appropriate due to the wind pattern and

was successful in explaining the changes in amplitude of the observed oscillations

Page 60: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

36 Chapter 2

in terms of the resonant interaction of the wind with the dominant natural mode of

the basin.

Degeneracy (several natural modes may have the same frequency) and

denseness (internal natural modes frequencies may vary over a continuous range)

(Münnich 1996) make the concept of natural modes fragile according to Maas and

Lam (1995), but we argue that the concept is still valid because the spatial

structure of the modes makes them clearly distinguishable and dictates which one

of them is excited by a wind with a given horizontal structure.

2.10 Conclusions

The azimuthal and vertical mode one Kelvin wave previously identified in

Lake Kinneret by Serruya (1975), Ou and Bennett (1979) and Antenucci et al.

(2000) was shown to travel around the entire basin and to possess a natural period

of 22.6 h, shorter than the observed periodicity of 24 h. The wave amplitude is

intensified at the southern and northern extremities of the lake where the radius of

curvature is smallest, in agreement with the numerical results of Schwab (1977)

for a flat-bottomed Lake Ontario and the analytical solution for a flat bottom

elliptical basin of Antenucci and Imberger (2001a).

The second dominant internal mode was made up of two counter-rotating

10.5-h-period Poincaré waves of the first vertical mode that meet the dispersion

relationships for circular basins. Once again this wave was also phase adjusted

each day by the wind forcing leading to a spectral peak close to 12 h. The basin

shape was shown to have two influences on the wave motion. First, the plan shape

determined, through the perimeter radius of curvature, the azimuthal Kelvin wave

amplitude variations, smaller radii leading to an amplification of the amplitude.

Page 61: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 37

By contrast, the effect on the 10.5-h period natural mode was to generate the two

counter-rotating Poincaré cells. Second, a sloping bottom induced an

amplification of the horizontal wave velocity over the slope similar to what has

previously been predicted in vertical two-dimensional basins; this is likely

produced by focusing of internal wave rays. In Lake Kinneret, gently sloping

regions have larger bottom velocities and thus higher potential for sediment

resuspension. We also showed that the elapsed time between the ends of two

consecutive wind events determined whether resonance between the wind and

wave motion occurred on a diurnal time scale.

2.11 Acknowledgments

The field experiment was funded by the Israeli Water Commission, via the

Kinneret modelling project, conducted jointly by the Kinneret Limnological

Laboratory (KLL) and the Centre for Water Research (CWR). Credits for the

success of the field campaign go to the Field Operations Group at CWR and to the

KLL laboratory. The authors thank the constructive and opportune comments by

Erich Bäuerle, Greg Ivey, Geoff Wake, Kevin Lamb, Roman Stocker, and two

reviewers. The first author gratefully acknowledges the support of an International

Postgraduate Research Scholarship and an ad-hoc CWR scholarship. This article

represents Centre for Water Research contribution ED 1671 AG.

Page 62: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

38 Chapter 2

Table 2.1. Details of the deployment

Station Deployment

period

Depth

(m)

Sampling

interval (s) Depth of thermistors

T4 01 Jan to 05 Jul 39.0 20 0.5 m and every 2 m from 2 m

Tg 22 Apr to 20 Jul 19.9 10 every 0.75 m from 0.75 m

Tf 22 Apr to 19 Jul 19.9 10 every 0.75 m from 0.75 m

T7 22 Apr to 20 Jul 20.3 10 every 0.75 m from 0.75 m

Tv 18 Jun to 02 Jul 29.5 10 every 0.75 m from 0.75 m

T9 17 Jun to 02 Jul 20.2 10 every 0.75 m from 0.75 m

Page 63: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 39

Table 2.2. Results of coherence and phase for 23°C isotherm vertical displacement for field data and ELCOM results. Station Tf is selected as

reference for zero phase when LDS or ELCOM results are analysed independently. When LDS and ELCOM results are compared, LDS data

are taken as reference.

LDS (compared with Tf) ELCOM (compared with Tf) LDS-ELCOM comparison

Station Coh2

24 h Phase 24 h (deg)

Coh2

12 h Phase 12 h (deg)

Coh2

24 h Phase 24 h (deg)

Coh2

12 h Phase 12 h (deg)

Coh2

24 h Phase 24 h (deg)

Coh2

12 h Phase 12 h (deg)

Tg 1.00 -55 0.47 91 0.94 -31 0.88 131 0.89 4 0.18 24 Tf 1.00 0 1.00 0 1.00 0 1.00 0 0.98 -19 0.99 -11 T9 1.00 114 0.94 -111 0.81 136 0.98 -98 0.72 6 0.89 3 T7 0.99 166 0.81 110 0.99 -151 0.95 -179 0.97 25 0.96 60 Tv 0.98 -11 0.98 14 1.00 -2 0.99 -6 0.96 -10 0.97 -31 T4 0.92 -40 0.93 -129 0.75 -179 0.92 103 0.72 -146 0.89 -134

Page 64: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

40 Chapter 2

Table 2.3. Estimates of dissipation time scale, damped natural period, and

resonant amplitude, obtained from the field data and from simulations with

different grid size. vwall is the simulation in a basin with flat bottom and the

horizontal shape of the perimeter of Lake Kinneret at the depth of the thermocline.

The resonant amplitude was non-dimensionalized by the resonant amplitude

estimated with the field parameters, BF.

Case 1/α (day) T (h) B/BF

Field 3.9 22.61 1

vwall 3.4 22.62 0.96

100×100 m 2.8 22.63 0.89

200×200 m 2.2 22.65 0.79

400×400 m 2.1 22.65 0.77

Page 65: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 41

0 2 4 6 8 10 12 140

2

4

6

8

10

12

14

16

18

20

22

Easting (km)

Nor

thin

g (k

m)

10

20

30

En Gev

Tabgha

Tf

Tg

T7

T9

T4

Tv

Figure 2.1. Bathymetry of Lake Kinneret. Filled and open dots indicate the

locations of LDS and external wind stations. The dashed contour indicates the

perimeter of the flat bottom basin used in one of the simulations. The origin of the

map coordinate system is situated at 35.51°N, 32.70°E.

Page 66: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

42 Chapter 2

0

5

10

Spee

d (m

s−

1 ) a

171 172 173 174 175 176 177 178 179 180 181 1820

90

180

270

360

Day in 2001

Dir

. (de

g)

b

10−6

10−5

10−4

10−3

102

104

106

Frequency (Hz)Spec

. den

s. (

m2 s

−2 H

z−1 )

24 h 12 h

c

Figure 2.2 (a) Wind speed and (b) wind direction (meteorological convention)

measured 1.5 m over the lake surface at station Tv. (c) Power spectral density of

the wind speed in (a).

Page 67: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 43

5

15

25

a) Tg

5

15

25

b) Tf

5

15

25

c) T9

5

15

25

d) T7

Dep

th (

m)

5

15

25

e) Tv

171 172 173 174 175 176 177 178 179 180 181 182

5

15

25

f) T4

Day in 2001

Figure 2.3. Record of isotherm displacements. In each panel, the isotherms are

26, 23, and 20 °C from top to bottom. Isotherm depths were determined through

linear interpolation of temperature records at fixed depths. (a) Tg, (b) Tf, (c) T9,

(d) T7, (e) Tv, and (f) T4.

Page 68: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

44 Chapter 2

10−6

10−5

10−4

10−3

100

105

1010

1015

1020

Frequency (Hz)

Spec

tral

den

sity

(m

2 Hz−

1 )Tg

Tf

T9

T7

Tv

T4

24 h 12 h

field

ELCOM

Figure 2.4. Power spectra of measured and simulated vertical displacement of the

23°C isotherm. Spectra have been smoothed in the frequency domain to improve

confidence. The two dotted lines near the bottom define confidence at the 95%

level. Vertical dotted lines mark periods of 24 and 12 h. Offset is 3 logarithmic

cycles.

Page 69: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 45

15 20 25 30

0

5

10

15

20

25

30

35

40

Temperature (°C)

Dep

th (

m)

a

0 0.006 0.012N (Hz)

b

Figure 2.5. (a) Mean temperature profile at station T4 for days 170.5 to 171.5 in

2001. (b) Associated buoyancy frequency.

Page 70: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

46 Chapter 2

1020304050

TgTfT9T7

a

Dep

th (

m)

−30

−20

−10

0 Tg

TfT9T7

b

170 171 172 173 174 175 176 177 178 179 180 181−15

−10

−5

0 Tg

Tf

T9T7

c

Day in 2001

Dis

plac

emen

t (m

)

Figure 2.6. (a) Depth of the 23°C isotherm displacements at stations along the 20

m isobath (with 10 m offset). Vertical displacement band-passed around (b) 24

hours (-10 m offset), and (c) 12 hours (-5 m offset). Dashed lines indicate the

propagation of the crests of the band-passed signals and are replaced by dotted

lines where the suggested propagation does not agree with an azimuthal mode one

wave propagating around the whole basin. The filled circles toward the bottom of

panels (b) and (c) are situated at times where there is a crest at station Tg and are

plotted to illustrate the phase lag between stations Tg and T7.

Page 71: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 47

−50

−40

−30

−20

−10

0 0 h

3 h

6 h

9 h

12 h

15 h

18 h

21 h

24 h

Tg Tf T9 T7 a

Dis

plac

emen

t (m

)

0 10 20−35

−30

−25

−20

−15

−10

−5

0

5

0 h

3 h

6 h

9 h

12 h

15 h

18 h

21 h

24 h

Tg Tf T9 T7 c

Distance (km)

Dis

plac

emen

t (m

)0 h

3 h

6 h

9 h

12 h

15 h

18 h

21 h

24 h

Tf Tv T4 b

0 2 4 6

0 h

3 h

6 h

9 h

12 h

15 h

18 h

21 h

24 h

Tf Tv T4 d

Distance (km)

Figure 2.7. Temporal evolution of the 24-h component of the 23°C isotherm

vertical displacement along the (a) azimuthal and (b) radial direction, and of the

12-h period component along the (c) azimuthal and (d) radial direction. Solid lines

suggest the isotherm displacements from equilibrium position (horizontal dash-dot

lines) from data available at the stations (vertical dashed lines) and are replaced by

dotted lines where the structure does not resemble an azimuthal mode one wave

propagating around the whole basin. The numbers on the right of each panel

indicate time from the first transect, which is at day 177.71 in panel (a) 177.82 in

panel (b), at day 177.98 in panel (c) and at day 177.81 in panel (d). Each isotherm

and its equilibrium position are offset by –6 m in panels (a) and (b) and by –4 m

in panels (c) and (d).

Page 72: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

48 Chapter 2

17

18

19

20

21

22

23

24

25

26

Isot

herm

(°C

)

0 0.5 1

a) Tf

−180 0 180

17

18

19

20

21

22

23

24

25

26

Isot

herm

(°C

)

d) Tf

Coherence squared 0 0.5 1

b) T9

−180 0 180 Phase (deg)

e) T9

0 0.5 1

c) Tv

−180 0 180

f) Tv

Figure 2.8. Profiles of coherence squared (solid line for field data and dashed line

for ELCOM results), and phase (circles for field data and triangles for ELCOM

results) of the vertical displacement of 24-h (a, b, and c) and 12-h (d, e, and f)

period oscillations. Coherence and phase are relative to the 23°C isotherm.

Page 73: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 49

Figu

re 2

.9.

(a-f

) D

epth

of

the

23°C

is

othe

rm

from

field

da

ta

and

ELC

OM

sim

ulat

ion;

(g

-l)

23°C

isot

herm

ve

rtica

l

disp

lace

men

ts

band

-pas

s

filte

red

arou

nd 2

4 h;

(m

-r)

verti

cal

disp

lace

men

ts

band

-pas

s fil

tere

d ar

ound

12 h

.

10 15 20

Tg

a

−505

g

−202

m

10 15 20

Tf

b

−505

h

−202

n

10 15 20

T9

Depth (m)

c

−505

i

−202

o

10 15 20

T7

d

−505

Vertical displacement (m)

j

−202

p

10 15 20

Tv

e

−505

k

−202

q

172

174

176

178

180

10 15 20

T4

f

172

174

176

178

180

−505

Day

in 2

001

l

172

174

176

178

180

−202

rfi

eld

EL

CO

M

Page 74: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

50 Chapter 2

Figure 2.10. Dominant natural modes of a basin with the horizontal shape of lake Kinneret 15 m isobath, flat bottom, and continuous stratification. (a) Power spectra of isotherm displacement and (b) cyclonic component of the rotary power spectra of isopycnal velocity for the 23-h-period natural mode; (c) power spectra of isotherm displacement and (d) difference between the anticyclonic and cyclonic components of the rotary power spectra of isopycnal velocity for the 10.5-h-period natural mode (white lines separate zones with opposite rotation). From top to bottom, the rows show results for the 26.5, 23, and 17°C isotherms, located in the epilimnion, metalimnion and hypolimnion respectively. The units of power spectra of vertical displacement are m2 Hz-1 and the units of rotary power spectra of isopycnal velocity are m2 s-2 Hz-1

Page 75: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 51

Figure 2.11 Dominant natural modes of lake Kinneret. (a) Power spectra of isotherm displacement and (b) cyclonic component of the rotary power spectra of isopycnal velocity for the 23-h-period natural mode; (c) power spectra of isotherm displacement and (d) difference between the anticyclonic and cyclonic components of the rotary power spectra of isopycnal velocity for the 10.5-h-period natural mode. From top to bottom, the rows show results for the 26.5, 23, and 17°C isotherms, located in the epilimnion, metalimnion, and hypolimnion respectively. The units of power spectra of vertical displacement are m2 Hz-1 and the units of rotary power spectra of isopycnal velocity are m2 s-2 Hz-1. The dotted and solid lines show the perimeter of the lake at the levels of the free surface and of the indicated isotherm

Page 76: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

52 Chapter 2

Ueast

at T/4N Sa

Usouth

at T/2N Sb

Ueast

at 3T/4N Sc

Usouth

at TN Sd

−0.04 −0.02 0.00 0.02 0.04 m s−1

Figure 2.12. Snapshots of the relevant component of the 10.5 h band-passed

velocity along the north-south axis of Lake Kinneret. (a) and (c) show velocity

perpendicular to the transect with positive values indicating velocities towards the

east; (b) and (d), velocity along the transect with positive values indicating

velocities towards the south. Solid white lines indicate zero velocity, dotted white

lines show the position of the 10.5-h band-passed 23°C isotherm oscillation, and

dashed white lines suggest the location of the boundary between the two counter-

rotating cells.

Page 77: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 53

−505

1015

spee

d (m

s−

1 )

a

5101520de

pth

(m)

Tf b

170 171 172 173 174 175 176 177 178

5101520de

pth

(m)

T7

Day in 2001

c

Figure 2.13. (a) West-east component of the wind at Tv and depth of the 23°C at

stations (b) Tf and (c) T7. Vertical dotted and dashed lines indicate, respectively,

the start and the end of wind events.

.

Page 78: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

54 Chapter 2

1234

Imp

(m2 s

−1 ) a

12

18

24

Hou

r of

day b

22

24

26

Δt (

h)

c

1

2

3

4

Am

plitu

de (

m)

dTgTfT7

170 175 180 185 190 195 200 205 6

12

18

Day in 2001

T p

eak

(h) e

Figure 2.14. (a) Momentum input of the daily wind events, (b) temporal location

of the daily wind events, (c) elapsed time between the ends of two consecutive

daily wind events (the dashed line indicates Δt = 22.6 h), (d) amplitude of the

Kelvin wave, and (e) time of arrival of the Kelvin wave crest. Amplitude and time

of arrival of the wave crest are obtained by band-pass filtering around 24 h the

23°C isotherm displacements .

Page 79: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 55

Figure 2.15. Lake Kinneret with measured forcing. (a) Power spectra of isotherm

displacement and (b) cyclonic component of the rotary power spectra of isopycnal

velocity at 24 h; c) power spectra of isotherm displacement and d) difference

between the anticyclonic and cyclonic components of the rotary power spectra of

isopycnal velocity at 12 h. From top to bottom, the rows show results for the 26.5,

23, and 17°C isotherms, located in the epilimnion, metalimnion, and hypolimnion

respectively. The units of power spectra of vertical displacement are m2 Hz-1 and

the units of rotary power spectra of isopycnal velocity are m2 s-2 Hz-1. The dotted

and solid lines show the perimeter of the lake at the levels of the free surface and

of the indicated isotherm.

Page 80: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

56 Chapter 2

0 0.2 0.4 0.6 0.8 1

12

14

16

18

20

22

24

26

28

30

Nondimensional RMS vertical displacement

Dep

th (

m)

24 h field12 h field24 h Elcom12 h Elcom

Figure 2.16. RMS vertical displacement profile at station Tv calculated from

measured and simulated temperature. Profiles are nondimensionalized by the

RMS displacement at the most bottom thermistor.

Page 81: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 57

a) T

/4

0.02

m s

−1

b) T

/2

0.02

m s

−1

c) 3

T/4

0.02

m s

−1

d) T

0.02

m s

−1

Figu

re 2

.17.

Sn

apsh

ots

of th

e 21

°C is

othe

rmal

vel

ocity

for

the

sim

ulat

ion

of th

e fla

t bot

tom

bas

in

with

the

horiz

onta

l sha

pe o

f the

15

m is

obat

h of

lake

Kin

nere

t and

con

tinuo

us s

tratif

icat

ion

afte

r an

initi

al ti

lt. V

eloc

ities

are

ban

d-pa

ss fi

ltere

d ar

ound

10.

5 ho

urs.

The

thic

k lin

e sh

ows

a no

n-cr

oss

flow

verti

cal s

urfa

ce a

nd c

ircle

s ind

icat

e th

e co

unte

r-ro

tatin

g ce

lls

Page 82: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

58 Chapter 2

0 1 20

1

2

3

4

5

| ω /

f |

S (ci r

0−1 f−1)

antic

yclon

ic (1

:1,r1

)

cyclo

nic (1

:1,r1

)

a

0.5 1 1.5 2 2.50.6

0.7

0.8

S

r

/r

b

acyc

cyc

acy

c

Figure 2.18. (a) Dispersion relationships of the two cells of the 10.5-h-period

natural mode of a basin with the shape of lake Kinneret at the level of the

thermocline, superimposed in the dispersion relationships derived by Antenucci

and Imberger (2001a). The circle corresponds to the bigger cell in the middle and

the diamond corresponds to the smaller cell in the south. (b) Dependence of the

ratio of radii of the anticyclonic and cyclonic cells on the Burger number of the

anticyclonic cell. The dot indicates the case of Lake Kinneret.

Page 83: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Spatial structure of basin-scale internal waves 59

a b

Figure 2.19. Shaded areas indicate where the ratio of the bottom and critical

slopes is between 1/1.5 and 1.5 for (a) the 22.6-h period Kelvin natural mode and

(b) the 10.5-h period natural mode. The slope of the wave rays was calculated

with Eqs. 2.8 and 2.9 with the value buoyancy frequency at the bottom taken from

Fig. 2.5b. Dotted contours indicate the intersection of the 26.5, 23, and 17°C

isotherms mean levels with the bottom of the lake.

Page 84: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

60 Chapter 2

Page 85: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

61

Chapter 3

Classes of high-frequency motions and their implications for

mixing: observations from Lake Constance

3.1 Abstract

Four different types of high-frequency motion were identified in Lake

Constance from temperature field data. First, periodic high frequency waves, with

periods between 4 and 13 min, were generated by shear instabilities in the surface

layer of the lake during strong wind periods leading to active mixing in the

surface layer. These waves did not contribute significantly to the horizontal

transport of mass and energy because they dissipated before being able to

propagate any significant distance. Second, waves with similar frequency were

generated in the metalimnion over the Mainau sill by the basin-scale wave

induced shear. Third, internal solitary waves with energy in Fourier components

with periods between 13 min and 1.2 h were generated by nonlinear processes at

the crest of steepened basin-scale waves. These waves propagated large distances

without producing significant mixing until they reached the sloping end of the

lake, where they dissipated producing mixing in the benthic boundary layer.

Fourth, the strong horizontal density gradient generated by a full upwelling event

developed in an internal bore that also carried energy in the 13-min to 1.2-h

spectral band. Due to the different layer thickness ahead, the bore behaved as a

buoyant surface plume in the north side of the main basin and as a non-breaking

undular bore in the south; after interaction with the Mainau sill, the bore evolved

into one of the breaking undular type. The internal bore contributed actively to

vertical mixing as it propagated through the lake interior.

Page 86: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

62 Chapter 3

3.2 Introduction

A thermal stratification of the water column supports internal oscillatory

motions covering a wide range of spatial and temporal scales (Imberger 1998).

Basin-scale internal waves, with spatial structure and period of oscillation linked

to the size and shape of the basin, generate coherent motions in the entire basin.

Waves of this scale are of two different types. First, internal gravity waves that are

the result of a balance between gravity, Coriolis, and inertia forces (Mortimer

1974). This type of waves generally dominates the spectra of isotherm

displacement in stratified lakes and is widely documented. If the Earth’s rotation

does not affect the internal motions in the lake, these waves are standing seiches;

otherwise, they are of the Kelvin and Poincaré types the modal structure of which

is determined by the basin shape. The second type of basin-scale internal waves,

the topographic waves, are an oscillation governed by the conservation of

potential vorticity; vertical vorticity stems are stretched when the water deepens

and contracted when it shallows leading to a returning force. Although

topographic waves can dominate the internal wave motions (Csanady 1976;

Stocker and Hutter 1987), they are generally difficult to observe in a thermistor

chain record due their small potential energy content and long period.

If the sampling period is short enough, internal waves with frequency

close to the maximum buoyancy frequency can be observed (see Antenucci and

Imberger 2001b for a review). These waves are associated with shear instabilities

generated by density currents (Hamblin 1977), metalimnetic jets (Boegman et al.

2003), basin-scale internal waves (Stevens 1999) or the wind stress applied at the

surface of the lake (Gómez-Giraldo et al. 2007).

Page 87: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 63

Solitary waves, with a frequency that is small compared to that of the

shear generated waves, but large compared to that of the basin scale waves, are

also often observed in lakes when nonlinear effects are important. The

nonlinearity may arise from the interaction of basin-scale waves with the

bathymetry (Thorpe et al. 1996; Saggio and Imberger 1998), large amplitude of

basin-scale waves (Thorpe et al. 1972; Boegman et al. 2005a), large horizontal

density gradients as those generated by full upwelling (Monismith 1986) or by

supercritical conditions that create an internal bore (Gan and Ingram 1992).

Several observations of the propagation of internal surges in long lakes have been

reported (Hunkins and Fliegel 1973; Farmer 1978; Mortimer and Horn 1982;

Wiegand and Carmack 1986). In most of the cases, they were assumed to be

generated by steepening of large amplitude basin scale waves but lack of data

precluded any conclusion about the generation mechanism.

All these internal waves are part of the energy flux path in stratified lakes

and influence directly the spatial and temporal distribution of many transport

processes. Hence, an understanding of the generation, propagation and dissipation

of the internal waves is necessary to understand water quality processes in lakes

and reservoirs. Most of the basin-scale internal waves are generated as a direct

response to wind stresses acting on the surface of the lake and tilting the

metalimnion (Mortimer 1974). These waves lose energy directly to mixing and

dissipation in the benthic boundary layer as they slosh over the bottom roughness

elements. They also transfer energy to the solitary-like waves generated by

nonlinear steepening. These solitary waves are observed to propagate over long

distances losing little energy as they propagate (Osborne and Burch 1980; Apel et

al. 1985; Boegman et al. 2003). Solitary like waves dissipate most of their energy

Page 88: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

64 Chapter 3

(up to a 90-95%) on the reflection off the sloping boundaries of the lake

(Helfrich1992; Michallet and Ivey 1999; Boegman et al. 2005b); the energy loss

contributes to sustain the benthic boundary layer turbulence.

Basin scale waves also lose energy directly through the formation of shear

instabilities. Small perturbations grow when the background shear is enough to

lower the gradient Richardson number below a critical threshold. The shear-

generated waves do not contribute significantly to horizontal transport of energy

because they dissipate over short distances (Boegman et al. 2003). However, they

are an important mechanism for vertical transport of energy and erosion of the

stratification (Gómez-Giraldo et al. 2007).

In Upper Lake Constance, basin-scale waves have been widely studied

(see Appt et al. 2004 for a review). The lack of high-resolution sampling has

prevented the observation of high frequency waves in the lake, but solitary waves

and shear generated waves were foreshadowed because of the large amplitude of

the basin-scale internal waves (Hutter et al. 1998) and the frequent strong wind

events. Further, even though it is well known (Laval 2003; Marti 2004) that the

wind is the main driver for the motions in a lake, the variability of the wind field

over Lake Constance had not been studied previously and the modelling of the

hydrodynamic motions has been based on simplifications and assumptions on the

wind field (e.g. Hutter et al. 1998; Wang et al. 2000).

A field campaign was conducted in the autumn of 2001 with two main

objectives. First, to measure the spatial variations in the wind field and relate the

basin-scale internal wave field and the dynamics of the lake to the wind field

(Appt et al. 2004). Second, to investigate and describe the high frequency internal

waves and relate these to the activity of basin-scale internal waves and to the wind

Page 89: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 65

forcing events. To achieve this, temperature was recorded at a sampling rate of 10

seconds at 8 locations throughout Lake Constance. The results from this work are

presented here.

3.3 Field site

Lake Constance (Fig. 3.1) is an Alpine Lake used for drinking water

supply, recreation, and commercial fishery. Dynamically, a shallow channel

separates Upper Lake Constance, the larger part of Lake Constance, and the

smaller and shallower Lower Lake Constance. The sill of Mainau, where the

depth at the talweg reduces to 100 m, divides Upper Lake Constance into the main

basin and Lake Überlingen. The main basin has maximum depth of 252 m, mean

depth of 101m, maximum width of 14 km, mean width of 9.3 km, and a length of

47 km. Lake Überlingen has maximum depth of 147 m, mean depth of 84 m,

mean width of 2.3 km, and a length of 16 km. Temperature is the dominant

parameter in the seasonal density stratification, with salinity being important only

when the temperature is close to 4°C (see Appt et al. 2004 for a review).

3.4 Field equipment

Eight Lake Diagnostic Systems (LDSs), each equipped with a 100-m-long

thermistor chain, wind speed, and direction sensor were installed in Upper Lake

Constance from 15 October to 17 November (days 288 to 321) of 2001. Five of

the stations were located in the main basin, one on the sill of Mainau and the

remaining two in Lake Überlingen (Fig. 3.1). Station s7 also had a full

meteorological suite of sensors measuring relative air humidity, air temperature,

incident short wave radiation, and total net radiation. The sampling rate was 10 or

Page 90: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

66 Chapter 3

30 s. Each thermistor chain had 51 thermistors with an accuracy of 0.01°C and a

resolution of 0.001°C spaced 0.75 m for the top 30 m and then at increasing

intervals from 1 m to 15 m at depth. A more detailed description of the equipment

and a general overview of the data are given by Appt and Stumpp (2002)

3.5 Methods

Traditional Fourier methods are commonly used for the identification of

dominant frequencies in temperature time series (Mortimer and Horn 1982;

Lemmin 1987). However, they are not effective in identifying periodic

components that appear intermittently in the signal, as is the case of high

frequency internal waves in the 33-day continuous temperature record. In

addition, phase information is lost with Fourier analysis. In order to keep the time

frequency information of the intermittent contributions and investigate their time

dependence to the wind forcing and to the phase of the basin-scale internal waves,

we carried out instead a wavelet transform analysis (Kaiser 1994).

The continuous wavelet transform of a discrete time series xn is defined as

its convolution with a scaled and translated version of a normalized wavelet

function ψ:

( ) ( )1

0

N*

n n'n'

n' n tW s x

ψ−

=

⎡ ⎤−= ⎢ ⎥

⎣ ⎦∑ , (3.1)

where N is the length of the series, δt is the sampling period, s is the wavelet

scale, and * indicates the complex conjugate (Torrence and Compo 1998). We

used a normalized Morlet wavelet:

( )2

0

1 21 4 2i t ttt e

sωδ eψ π − −⎛ ⎞= ⎜ ⎟

⎝ ⎠ (3.2)

Page 91: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 67

where t is time and ωo is the nondimensional frequency. The scale s is related to

the Fourier frequency f by:

2

0 0

4

2

π=

ω + + ω

sf (3.3)

(Torrence and Compo 1998). The wavelet power spectrum is defined as

( ) 2nW s (Torrence and Compo 1998).

Fourier or wavelet transforms are traditionally applied to an isotherm

displacements time series. The selection of the isotherm is subjective, but the

isotherm in the middle of the thermocline is usually chosen. During the field

campaign in Lake Constance, the relative position of the isotherms in the water

column changed due the strong cooling experienced over the measurement period.

Hence, instead of selecting an isotherm, we analysed the time frequency contents

of the integrated potential energy, defined as

( ) ( )2

1

H

H

IPE t z,t g z dzρ= ∫ , (3.4)

(Antenucci et al. 2000) where ρ is density, g is the gravitational acceleration and z

the vertical coordinate increasing upwards. For integration over the entire water

column H1 and H2 should be the elevations of the bottom and surface respectively.

However, the LDSs only measured the top 100 m of the water column; hence the

datum was set at the elevation of the most bottom thermistor, so H1 = 0 m and H2

= 100 m.

The information extracted from the field data was complemented with

numerical simulations with the Estuary and Lake Computer Model (ELCOM) to

investigate the propagation of a front developed after a full upwelling event. A

Page 92: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

68 Chapter 3

description of the model is presented by Hodges et al. (2000) and the specific

validation for the basin-scale dynamics of Lake Constance was carried out by

Appt et al. (2004). We used a hydrostatic version of the model and a relatively

coarse horizontal grid (400 m x 400 m) so details of the evolution of nonlinear

phenomena with large vertical velocities, such as the propagation of a density

front (see below), could not be studied in detail. However, Appt et al. (2004)

showed that the model reproduces well the time of arriving of the front to the LDS

locations.

3.6 Background stratification, forcing events, and basin-scale motions

A complete description of the evolution of the stratification and a

characterization of the basin-scale internal waves is given by Appt et al. (2004).

Here we only present those aspects that are important to define the background

conditions on which the high frequency internal waves were generated.

During the experiment, the lake experienced a typical autumn cooling.

Between days 288 and 310, the wind events were short and had maximum speeds

ranging between 4 and 9 m s-1 (3.2a). During this period, the surface temperature

dropped gradually from 15 to 12 °C (Fig. 3.2b). The thick metalimnion was

located between 10 and 40 m deep with the maximum buoyancy frequency

located at the top of the metalimnion and with a value close to 5.5×10-3 Hz (Fig.

3.2c). A strong WSW wind of long duration, starting on day 310, resulted in a

surface temperature drop to 10°C in only 3 days. The strong wind also caused

deep mixing probably both by overturning and convective cooling of the water

column and the buoyancy frequency reduced between days 313 and 315 to a value

close to 1×10-3 Hz. After this, the surface layer thickness was about 25 m and the

Page 93: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 69

metalimnion was thinner with a weaker stratification than during the initial

monitoring period.

The dominant basin-scale internal wave was a vertical and azimuthal mode

one Kelvin wave with a period around 90 h that propagated around the entire lake

including Lake Überlingen (Bäuerle et al. 1998; Appt 2004). Poincaré waves were

also part of the basin-scale internal wave field, but these waves had different

dominant periods in different parts of the basin depending on the local width of

the basin (Wang et al. 2000; Appt et al. 2004).

3.7 High frequency waves. Prominent features

We calculated the wavelet power spectra of the IPE to identify the

dominant frequencies and plotted it together with the wind and the IPE to identify

the most likely generation mechanism. Fig. 3.3 shows the results for station s2 for

periods smaller than 6 hours; a log2 scale is used because it provides a good power

resolution. Two features are evident; first, an increased power in the 4- to 13-min

period band during periods of strong wind and second, an increase in power in the

entire 13-min to 1.2-h period band after the crests of the IPE signal reached a

magnitude larger than 4.826x107 kg m-2 on days 306.65 and 314.49. Weaker

bursts of energy, covering a smaller period subrange, were observed in this period

band in association with smaller crests of the IPE signal (e.g. days 294.7 and 310).

This association with the crests of the IPE signal, which is a direct measure of the

vertical mode 1 basin-scale internal waves amplitude, suggests that the features in

this frequency band were generated by nonlinear processes. However, the two

more energetic high frequency events observed at s2 on days 306.65 and 314.49

had different generation mechanisms and different characteristics (see below). In

Page 94: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

70 Chapter 3

what follows, we describe the waves in the 4- to 13-min period band and the two

dominant events in the 13-min to 1.2-h period band, that we will call events I and

II.

3.8 Waves with periods between 4 and 13 minutes

Waves in this frequency band were excited whenever there was a strong

wind event (days 294, 304.8, 313, and 315) indicating that they were generated by

instabilities induced by the strong shear produced by the wind. On days 296 and

300 the shear was provided by the vertical mode one Kelvin wave at the location

of wave trough, which generates strong shear in the metalimnion (Antenucci et al.

2000).

Fig. 3.4a shows the waves with periods of minutes generated at s5 by the

wind event of day 304.8. The waves were of the first vertical mode and the

vertical displacements were larger in the base of the surface layer than in the

stratified region. This suggests that these high frequency internal waves were

generated by the shear induced by the wind in the surface layer similar to that

observed at the crests of Kelvin waves in Lake Kinneret during strong wind

conditions (Gómez-Giraldo et al. 2007). Fig. 3.4b shows the high frequency

waves generated at the trough of the basin-scale waves at s3 on day 300. These

waves were also of the first vertical mode and were observed throughout the

metalimnion with larger amplitudes in the region of stronger stratification, i.e. at

about 26 m depth. This is in agreement with waves generated by shear instabilities

centred in the metalimnion (Stevens 1999).

Time series of log2 of the wavelet power in the band 4- to 13-min were

compared to those of wind speed and IPE for several stations around the lake,

Page 95: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 71

with the purpose of identifying direct temporal and spatial dependencies. Fig. 3.5

shows the data for day 304.8 and it is seen that the energy rose simultaneously at

all stations through the entire lake in direct response to the wind forcing. The

amount of energy fed into these waves varied depending on the local wind

strength and on the relative phase of the wind and the basin-scale internal waves,

as discussed by Antenucci and Imberger (2001b) and Gómez-Giraldo et al.

(2007), but there is not a direct association of the generation of the high-frequency

waves with any particular characteristic of the IPE signature. The waves

generated by high shear associated with the basin-scale waves were localized and

were only very energetic at s3 (Fig. 3.6), noticeable at s2 (Fig. 3.3c) and absent at

the other stations. The wind was weak even at s3 (Fig. 3.6), indicating that the

generation of these waves were not directly associated with the wind. The IPE

signal (Fig. 3.6) did not present any feature suggesting a generation mechanism

either. The fact that these waves were localized in an area around the sill of

Mainau, points to a generation mechanism associated to a local effect of the

bathymetry. It is likely that the shear induced by basin-scale internal waves was

largest over the sill due to the local acceleration, so we used the velocity field

calculated with ELCOM to explore this idea. The model was started on day 292

with initial conditions, forcing and model parameters as described in Appt et al.

(2004). Fig 3.7a shows the comparison between the model and field isotherm

displacement at station s3 around the time the discussed waves were observed;

clearly ELCOM reproduced well the basin scale motions and thus provided a

meaningful estimate of the velocity field induced by the Kelvin waves. To

evaluate whether the waves were generated by shear, we calculated the gradient

Richardson number (Rig) from the model results and applied the correction

Page 96: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

72 Chapter 3

derived from Simanjuntak, M. A. (unpub.) to take in account the effect of a coarse

scale being used for this estimation. Fig. 3.7b shows that, at s3, Rig dropped below

the critical value of 0.25 around a depth of 26 m just before the trough of the

Kelvin wave on day 300, making possible shear instabilities to develop. Strong

activity of waves in this frequency range was also observed at s3 during the

trough of the Kelvin wave well before the wind event on day 304.8 (Fig. 3.5) and

was probably also generated by wave associated shear.

There was no increase in the other high frequency bands at the time these

shear waves were observed, indicating that the energy was introduced directly at

this scale without cascading through the internal wave spectra.

3.9 High frequency event I

As seen in Fig. 3.3c, the wavelet estimate of the power spectral density

was elevated at periods between 13 min and 1.2 h over the period between 306.5

and 308 and was associated with an elevated IPE signal due to the large amplitude

of the basin-scale internal Kelvin wave that resulted after the strong wind event on

day 304. High frequency wave events like this one were not present before this

period at the stations located in Lake Überlingen or the western end of the main

basin, nor before day 313 for the other stations due to the lack of strong wind

events. Station s10 was the only location that showed energy, although moderate,

in this band (13 min to 1.2 h) from the beginning of the experiment. This was

probably due to waves generated by the degeneration of the Poincaré waves that

had a larger amplitude at this location (Appt et al. 2004).

Fig. 3.8 shows the isotherm contours for this event as observed at s3. It

reveals a leading vertical mode one solitary wave of depression (A) followed by a

Page 97: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 73

smaller vertical mode one solitary wave (B) and a second vertical mode nonlinear

wave (C). Fig. 3.9 shows the log2 of the wavelet power in the 13-min to 1.2-h

band, together with the wind speed and the IPE for the stations in Lake

Überlingen and the west end of the main basin. Clearly seen from this figure is the

propagation of this train of waves from s5 to the end of Lake Überlingen at s1 and

then propagating back after reflecting from the western boundary of the lake. The

propagation was independent of the wind and, being vertical mode 1 waves, their

speed was similar to that of the basin scale Kelvin wave and so it is not surprising

that the solitary waves remained almost phase locked with the crest of the Kelvin

wave. Fig. 3.10 shows the details of these high frequency waves and their wavelet

power in the 4- to 13-min and 13-min to 1.2-h bands. The larger amplitudes of the

leading solitary wave and the formation of an oscillatory tail suggest that the wave

train gained energy while it travelled from s3 to s2 (Fig. 3.10a-b) probably due to

further steepening of the basin scale wave between both stations (Horn et al.

2000) (note that the amplitude of the basin-scale wave at s2 was larger than at s3).

Note also that, due to a larger phase speed resulting from its larger amplitude, the

leading solitary wave separated from the second solitary wave. The vertical mode

two solitary wave, which is much slower, had been left behind and was not

observed at s2. After s2 the leading solitary wave lost energy to the trailing waves

as it propagated to s1 (Fig. 3.10b-c). This is the normal evolution of such a train

of nonlinear waves (Horn et al. 2000). Most of the energy in the leading solitary

wave and the trailing waves was lost in the reflection at the western end of the

lake (Fig. 3.10d). A similar observation of high frequency waves following a

steep front being dissipated after reflection at the end of Loch Ness was reported

by Thorpe et al. (1972). The observed reduction of amplitude from Fig. 3.10c to

Page 98: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

74 Chapter 3

Fig. 3.10d was about 80%, which is similar to estimates obtained from the

laboratory studies of Boegman et al. (2005b). The reflected solitary wave and

trailing waves gained energy again, probably from the basin-scale Kelvin wave,

on propagating from s1 to s2 and lost it again in the propagation to s3 (Fig. 3.10d-

f). A gap in the data at s5 prevents the investigation of propagation in that

direction but the fact that the solitary waves were not observed at s6 (Fig. 3.9)

indicates that their amplitude decreased drastically as they left Lake Überlingen

and moved into the wider main basin. This is similar to that observed in Babine

Lake by Farmer (1978).

From the peaks in wavelet power in Fig. 3.9, the observed speed of

propagation of the solitary wave between s3 and s2 was 0.27 m s-1. A simplified

weakly nonlinear, 2-layer analytical model of internal solitary waves (Benney

1966) allows estimating the phase speed of waves of finite amplitude as

013

c c aα= + (3.5)

where 2 1 1 20

2 1 2

h hc gh h

ρ − ρ=

ρ +is the linear long-wave phase speed, a is the

amplitude of the solitary wave, and 1 20

1 2

32

h hch h

−α = is the nonlinear coefficient.

With top and bottom layer thicknesses, h1 and h2, of 15 and 85 m, mean

temperatures of 12 and 7 °C ( 1ρ and 2ρ are the associated densities), and a = 7

m, c is 0.28 m s-1, in good agreement with the observations. This is 12% faster

than the linear phase speed of 0.25 m s-1.

Closer inspection of Fig. 3.8 reveals that the passing of the solitary waves

did not change significantly the separation between isotherms and did not change

Page 99: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 75

the elevation of the isotherms in addition to the drop produced by the basin-scale

waves. This indicates that the solitary waves did not create important mixing (and

dissipation) in its propagation.

Although the solitary waves were most frequently observed on the crest of

the internal Kelvin wave in Lake Überlingen, they can also evolve from

steepening of the basin-scale internal Poincaré waves in the main basin. Fig. 3.11

shows the formation and evolution of a solitary wave as it propagated from s5 to

s7, following the crest of the large amplitude Poincaré wave excited by the strong

wind event of day 304 (Appt et al. 2004).

3.10 High frequency event II

This high frequency event followed a full upwelling event east of the sill

of Mainau. A strong WSW wind event, with speeds reaching 12 m s-1 at 2.4 m

height that started early on day 313 and continued for almost three days, created

full upwelling in the north side of the main basin and east of the sill of Mainau.

Appt et al. (2004) explained the upwelling close to the sill in terms of the water

flow divergence created by the variable wind field and the difference in width of

both sub-basins, but did not explain the upwelling in the north side of the main

basin.

For an elongated basin such as Lake Constance, a dominant longitudinal

component of the wind sets up a longitudinal pressure gradient and the Coriolis

force creates a transverse force that, in turn, is balanced by a transverse pressure

gradient. A strong wind blowing along the major axes will produce upwelling at

the upwind end of the lake or along the side on the left of the wind (in the

northern hemisphere) depending on the ratio of the time scales that it takes to

Page 100: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

76 Chapter 3

produce the maximum longitudinal tilt and the time it takes for the transverse

gradient to develop:

24

t

i

t BU WT L

= = S (3.6)

where tt is the time required for the Coriolis force to produce lateral upwelling due

to the transversal tilt of the stratification and Ti is the period of the gravest

longitudinal seiche. Ti/4 is then the time that a constant wind takes to produce the

maximum amplitude of the seiche (Spigel and Imberger 1980). W is the

Wedderburn number, S is the Burger number, and L and B are the length and

width of the basin. The derivation of U is presented in the appendix.

If U is less than one, the upwelling occurs first along the long side of the

lake at the left side of the wind (northern hemisphere); if it is larger than one,

upwelling occurs first at the upwind end of the lake. The transverse component of

the wind modifies only slightly the distance from the lateral shore to the front of

the upwelling (Csanady 1977) and can be neglected. For the specific wind episode

that is considered below, Appt el al. (2004) showed that the transversal

acceleration due to the transverse wind was weaker than that due to the Coriolis

force generated by the longitudinal wind.

To estimate U, we approximated the thermal structure of Lake Constance

prior to the strong wind event by two layers. The thickness and mean temperature

were 20 m and 11.5°C for the top layer, and 80 m and 7°C for the bottom one.

The length and mean width of the main basin (37 km and 9.3 km) were used for L

and B respectively and f was set to 1.075x10-4 rad s-1. U was then 0.15, so

upwelling should have developed first along the north side of the basin. This was

effectively the situation described by Appt. et al. (2004) and reproduced by

Page 101: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 77

ELCOM simulations (Fig. 3.12a). The upwelling structure was modified by the

local bathymetry and the upwelling was intensified in the two areas to the right of

the shorelines with the same orientation of the wind, probably due to a stronger

offshore Ekman transport. Fig 3.12b shows the evolution of the upwelling under

an almost constant SW wind. With time, the upwelling front near the north shore

joined the upwelling produced in the southwest corner of the main basin. As the

wind changed its direction to NW, the offshore Ekman transport from the shore

next to s5 intensified and strong upwelling occurred in this area (Fig. 3.12c)

despite the reduced wind speed. Then the wind became north-easterly and the

Ekman transport was toward the north-west (Fig. 3.12d). The relaxation of the

upwelling, following the weakening and change of direction of the wind on day

313.7, did not show the cyclonic propagation characteristic of a Kelvin wave; the

front of the warm water in the main basin moved toward the west end of the basin,

while the front on the side of Lake Überlingen remained stationary (Figs. 3.12e,f).

Csanady (1977) reported a similar observation in Lake Ontario and explained it,

following Bennett (1973), in terms of the advection and phase speed of the waves

being in the same (opposite) directions in the north (south) sides of the lake. In the

case of the observations in Lake Constance, another plausible explanation is that

the propagation of the front was forced by the wind as suggested by the horizontal

convergence in surface transport observed in Figs. 3.12d-f.

At s5 (Fig. 3.12 e), in the northern shore of the main basin, the warm water

front propagated westward as a buoyant surface plume with a large head that

induced isotherm drops as large as 90 m (Fig. 3.13 a). The large separation of the

isotherms between 5 and 50 m depth and the abundant statically unstable

stratification patches in the water column (Figs. 3.13a, 3.14a,b) indicate that

Page 102: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

78 Chapter 3

mixing was active. Strong shear in the leading edge of the front generated large

billows like the one near the surface in day 313.745 and the one at mid-depth

(from 20 to 65 m deep) at day 313.752 (Fig. 3.14a), and the turbulent core

produced the large-scale mixing structure in Fig. 3.14b. These are all features of a

buoyant surface plume like that described by Luketina and Imberger (1989).

Before the passage of the plume, the entire depth was made of hypolimnetic

water; after the passage of the 90 m deep plume head, the metalimnion was

centred at around 40 m deep (Fig. 3.13a). High frequency waves were not evident

in or behind the plume.

The plume weakened in its propagation due to active mixing and, at s3, the

isotherms drop was only about 30 m and less mixing was perceived (Figs. 3.13b

and 3.14c). At s2, downstream of the sill, the front developed into an internal

undular bore with an isolated deep excursion of the isotherms at the front followed

by breaking nonlinear waves with a dominant period of about 19 min (0.013 day)

and overturning features above the troughs (Figs. 3.13c and 3.14d). Note that I

call it an undular bore if there is a two layer stratification ahead of the

perturbation, and call it a surface buoyant plume if the water column is relatively

well mixed ahead of the perturbation. Similar large collapsing undular bores have

been observed in Knight Inlet (Farmer and Smith 1980). The bore in Lake

Überlingen was followed by a long wake with abundant oscillations with vertical

excursions of up to 15 m and with clear signals of shear instabilities. Fig. 3.14e,

for instance, shows a group of billows in different stages of growth and

overturning. The period of the oscillations in this part of the wake following the

front was about 14 min (0.01 day). At s1, the wake had lost most of its energy and

only a couple of solitary-like waves remained (Fig. 3.14f). The frontal face of the

Page 103: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 79

leading waves had become more gently sloping and the rear front had steepened in

the shoaling process to the point where a convective instability occurred. This

behaviour was predicted by the numerical investigations of Vlasenko and Hutter

(2002), and the laboratory experiments of Michallet and Ivey (1999). In addition

to the generation of the bore downstream of the sill, the interaction of the front

with the sill also produced reflection of part of the energy of the incident front. A

secondary isotherm drop was observed at s5 on day 314.33 (Fig. 3.13e), 0.6 days

after the passage of the incident front.

In the southern side of the main basin, the westward propagating front had

a different behaviour to that observed along the north shore. The front reached s6

around day 314 (Fig. 3.12f) as a nonbreaking internal undular bore approximately

20 m high and that extended until day 314.15 (Fig. 3.15a). This undular bore

probably reflected from the surrounding sloping bottom and appear again at s6 on

day 314.28 (Fig. 3.15b). The isotherm behaviour indicates that weak mixing and

dissipation occurred within the bore, but the reduced amplitude of the reflected

bore suggests that it lost a significant amount of energy on shoaling and reflection

on the sloping bottom. The reflected bore was not identified in other LDSs but

whether it dissipated before reaching them or simply followed a path that did not

traverse them is an open question.

The strong wind event responsible for the full upwelling in the west of the

main basin also created large downwelling at the south-eastern end of the lake

(measured at s10, not shown) and energized large amplitude Poincaré waves

(Appt et al. 2004). Energy in the frequency band of the solitary waves was also

detected in the IPE wavelet analysis while the amplitude of the Poincaré waves

Page 104: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

80 Chapter 3

was large, but a picture of the expected solitary waves was unclear due to the

large spacing of the thermistors around 60 m, the thermocline depth at that time.

3.11 Discussion

We presented field observations from Lake Constance showing three

different processes that put energy in the high frequency band: periodic waves

generated by shear instability (either in the mixed layer or due to baroclinic shear

in the metalimnion), nonlinear steepening of large amplitude basin-scale internal

waves, and propagation of internal bores. Our observations provide evidence of

the occurrence, in a stratified lake, of several processes studied in the laboratory

or through numerical analysis, and provide valuable information in order to

understand the energy flux path in stratified lakes.

3.11.1 High-frequency waves generated by shear instability

Results of previous investigations of wind induced shear instabilities

(Gómez-Giraldo et al. 2007) indicate that part of the energy that they extract from

the background flow is converted to potential energy through mixing following

the billowing, and part of the energy can be transported vertically to a different

level where the local conditions are such that the energy can be returned to the

background flow or where a secondary instability can be triggered. Waves

generated by this mechanism dissipate before they can travel any significant

horizontal distance (Boegman et al. 2003; Gómez-Giraldo et al. 2007) so that their

action is confined to local dissipation and vertical mixing.

It was shown that shear instabilities in Lake Constance were also

generated by the baroclinic shear associated with intensification, over the Mainau

Page 105: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 81

sill, of the vertical shear of the vertical mode 1 basin-scale internal waves. Waves

generated by interfacial shear have been observed in laboratory experiments

(Thorpe 1968), in the ocean (Woods 1968), and stratified lakes (Stevens 1999).

These waves also dissipate quickly and only contribute to localized vertical

mixing.

3.11.2 Solitary waves

To verify that the solitary waves observed on day 306 resulted from

nonlinear degeneration of the large amplitude basin-scale waves, we used the

regime diagram of Horn et al. (2001). Simplifying the lake as a rectangular 150-m

deep basin with the 9°C defining a two-layer stratification, the top layer is 20 m

thick and the amplitude of the basin-scale wave on day 306 is 8 m. Hence, the

values for amplitude ratio and the depth ratio are 0.4 and 0.13, respectively, and

the regime diagram of Horn et al. (2001) predicts that the basin-scale waves are

degenerated into solitary waves by nonlinear steepening. The sequence of vertical

modes one and two solitary waves observed over the sill is generated by fission of

nonlinear waves due to their interaction with topography and has been studied in

the laboratory by Helfrich and Melville (1986) and numerically by Lamb (2002)

and Vlasenko and Hutter (2002). The investigations of the later authors were in

fact completed following three hypothetical paths over the sill of Mainau. Station

s3 is in the path of a trajectory where they predicted transformation without

disintegration of the solitary wave, but a different trajectory of the solitary wave

before reaching s3 to that assumed by Vlasenko and Hutter (2002), different

background conditions or different characteristics of the incident solitary wave

may have produced the regime shift. It is likely that the topographic changes at the

Page 106: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

82 Chapter 3

sill also participated in the nonlinear evolution and transformation of the solitary

waves.

The solitary waves generated by nonlinear steepening and modified after

interaction with the sill only change the stratification slightly as it passes through

(see Fig. 3.8), indicating that this type of waves generate little dissipation and

vertical mixing in the interior of the lake. Similar conclusions can be drawn from

observations in other lakes (Thorpe et al. 1972; Hunkins and Fliegel 1973; Farmer

1978; Wiegand and Carmack 1986) and in the ocean (Osborne and Burch 1980;

Apel et al. 1985). In contrast, reflection at the boundaries causes lost of most of

the energy in these waves. Significant loss of energy in solitary waves was also

observed in Lake Pusiano by Boegman et al. (2005b) and the laboratory

experiments of Helfrich (1992), Michallet and Ivey (1999) and Boegman et al.

(2005b). The length scale of the small incident solitary waves observed at s1 is

400 m and the length of the slope is about 5 km. For their small ratio, 0.08, the

results of Michallet and Ivey (1995) predict the energy in the reflected waves to

be as small as 10% of the energy in the incident waves, which is in agreement

with our observations of the small energy in the reflected wave train. Similarly,

the results of Boegman et al. (2005b) also predict reflection of only about 20% of

the energy in the incident wave.

3.11.3 Internal bore

A highly nonlinear density distribution like that generated by the full

upwelling event is inefficient for formation of basin-scale internal waves

(Monismith 1986) and the basin-scale waves excited by this wind event were

smaller than those excited by weaker wind events (see Appt et al. 2004). Our data

Page 107: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 83

showed that, instead, it favoured the formation of an internal bore as predicted by

the regime diagram of Horn et al. (2001) when the amplitude of the initial tilt of

the stratification is larger than the upper layer thickness, as is the case for full

upwelling. The bore, however, had a different structure in the north (s5) and south

(s6) shores of the main basin and in Lake Überlingen (s2) that follows the regime

classification for internal bores defined experimentally by Rottman and Simpson

(1989) in terms of the strength of the bore, which is the ratio of the upstream (hu)

to the downstream depth (hd) of the interface. Table 3.1 presents the surface layer

thickness upstream and downstream of the bore, as observed at the three stations

where the different types of bore were observed, the bore strength and the

expected type of bore. According to Rottman and Simpson (1989) classification, a

bore is undular and laminar for hu/hd < 2.3, is undular with turbulent mixing

behind the crests for 2.3 < hu/hd < 3.5, and is a buoyant surface plume (or surface

gravity current) for hu/hd > 3.5, which agrees well with our observations. The

speed of the bore (U) and of the dissipation that it produces (E) can be estimated

using the formulas derived by Klemp et al. (1997) based on two-layer hydraulics:

( )( )

( )2

2

2

3u u u d

u u a u d u

h H h H h hUg h H h H h H h h h

− − −=

′ + + − (3.7)

( )( )(( ) ( )

)33

22

2

2 2d u u

u u u d

U H H h H h h hE

h H h H h h

ρ − − −=

− − −d (3.8)

where ( )2 1g g 2ρ ρ ρ′ = − is the reduced gravity, g is gravity, ρ is the mean

density and H is the total flow depth. Despite controversy about the assumptions

in the derivation, these formulas produce good estimates when the thermocline

Page 108: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

84 Chapter 3

depth is much smaller than the total depth. A crude estimation of the buoyancy

flux (B) can be obtained by assuming a Richardson flux number

fBRi

B E=

+ (3.9)

of 0.20 as is usual in geophysical flows. The estimates of U, E, and B are also

presented in Table 3.1 together with direct estimates of speed and buoyancy flux

from field data (Uobs and BBobs). Uobs was estimated from the time of arrival of the

bore at different stations and the distances between them; at s5, BobsB

)

was estimated

as

(3.10) ( 313 9 414 1obs . . obsE IPE IPE U= −

where the subindex of IPE indicates the time of measurement. At day 313.9 the

mixed roller inside the head of the buoyant surface plume is passing s5; at day

314.1 a profile well behind the roller, hence unaffected by the strong mixing,

crosses s5 (Fig. 3.13a). The good agreement with our data supports the

applicability of the classification of Rottman and Simpson (1989) and the

analytical results of Klemp et al. (1997). It is interesting to note that the difference

in the type of bore observed at s5 and s6 is an indirect effect of the Coriolis force

that generated different downstream interface depths.

There are several differences between the mechanisms involved in the

generation of the breaking undular bore downstream of the sill with those lee

waves observed in the atmosphere (e.g. Klemp and Lilly 1975) and in oceanic

inlet sills (e.g. Farmer and Smith 1980). First, the frontal flow is forced over the

sill by the baroclinic pressure gradient and the velocity profile is likely to have a

vertical structure that contributes to shear in the thermocline. Second, the density

structure was already highly nonlinear before flowing over the sill. Third, the flow

Page 109: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 85

changed quickly with the passage of the flow and the propagation of the front

generated a propagating bore instead of a steady hydraulic jump downstream of

the sill. Farmer and Smith (1980) showed that the qualitative response to the

interaction with the sill is affected by the density structure and phase and strength

of the tidal cycle. Such sensitivity to the background conditions is supported by

the numerical investigations of Afanasyev and Peltier (2001) and Lamb (2004).

Although our observations are qualitatively similar to those in the ocean and the

atmosphere, it is likely that the differences in the generation can also alter the type

of response, but further research is required. Lack of spatial resolution precluded a

detailed description and understanding of the flow generated by the interaction of

the front with the sill in Lake Constance but it is clear that it generated abundant

mixing over the entire thermocline in Lake Überlingen.

Field data (Farmer and Smith 1980) and numeric investigations (Lamb

2004) have shown that separation of the boundary layer may occur downstream of

the tip of the sill with coherent structures observed along the lee side. This

phenomenon could also happen and produce mixing and resuspension along the

bottom of Lake Überlingen but further research is needed to address this issue as

our thermistor chains covered only the upper 100 m of the water column.

It is a generalized idea that most of the mixing in the metalimnion of

stratified lakes occurs in the benthic boundary layer (Imberger 1998), where the

vertical mixing rates are at least one order of magnitude higher than in the interior

of the lake (Goudsmit et al. 1997; MacIntyre et al. 1999). The mixing in the

benthic boundary layer is generated by shoaling (Taylor 1993), shear-generated

dissipation (Lemckert and Imberger 1998), and geometric focusing (Maas and

Lam 1995; Gómez-Giraldo et al. 2006) of basin-scale internal waves, and by

Page 110: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

86 Chapter 3

breaking of high frequency internal waves at reflection on the boundaries (De

Silva et al. 1997; Michallet and Ivey 1999). We showed evidence of such

transport of energy to the benthic boundary layer by solitary waves, but the strong

and quick mixing generated in the interior of the lake by the propagation of a front

poses the question whether events of this type, although relatively infrequent, may

have a more relevant influence in the evolution of the stratification and, as a

consequence, in the ecology of stratified lakes.

Beside the indirect effect on the ecology due to changes in stratification,

solitary waves like those in event I may also play a direct role in the horizontal

distribution of species in the lake as they can transport particles during

propagation (Lamb 1997). Pineda (1999) showed that internal bores accumulate

larvae and very likely transport it for long distances, so it is probable that

propagation of fronts also contribute directly to the horizontal distribution of

species in Lake Constance.

3.12 Conclusions

Three types of high-frequency motions are observed to contribute in

different ways to the energy cascade from large to small scales in stratified lakes.

First, high-frequency internal waves generated by shear instabilities dissipate

quickly producing localized vertical mixing. Second, solitary waves generated by

nonlinear steepening of basin-scale internal waves can travel long distances

without losing energy significantly until they break at the boundaries, producing

intensive mixing in the benthic boundary layer. Internal bores produce strong

mixing as they propagate through the lake, contributing to sporadic mixing in the

interior of the lake.

Page 111: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 87

3.13 Acknowledgements

The field experiment was funded by the Deutsche

Forschungsgemeinschaft (DFG), via the Modelling of high-frequency internal

waves in Lakes project, and was conducted by the Field Operations group at the

Centre for Water Research and the Institut für Wasserbau at the University of

Stuttgart. In particular, we thank Prof. Helmut Kobus for the coordination of the

project, and Sheree Feaver, Jochen Appt and Simone Moedinger for their

organization of the field campaign. The wavelet software was provided by C.

Torrence and G. Compo. The first author acknowledges the support of an

International Postgraduate Research Scholarship, a UWA Scholarship for

International Research Fees, and an ad-hoc CWR scholarship. This paper

represents Centre for Water Research reference ED 1673 AG.

Appendix. Derivation of U

For elongated basins where the dynamics are not affected by the Earth’s

rotation, a longitudinal wind sets up a longitudinal tilt of the stratification. If the

wind is strong enough the magnitude of the tilt is such that hypolimnetic reaches

the surface at the upwind end of the lake. If the Earth’s rotation affects the

dynamics, a transversal tilt of the stratification is produced by the Coriolis force

and could lead to upwelling along the long side of the lake at the left of the wind

(in the northern hemisphere). Assuming that upwelling occurs, and using scaling

arguments, we derive below a criteria to estimate if upwelling happens first at the

upwind end or along the long side of the lake.

Page 112: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

88 Chapter 3

Assume a rectangular lake with length L and width B and a two-layer stratification

with density and velocity 1ρ and for the top layer, and 1u 2ρ and for the

bottom layer.

2u

In the transversal direction, the Coriolis force balances the transversal pressure

gradient balance and we get

( )1 2 1 11

0

2 21 h g hu f gB B

ρ ρρ

−′= = (3.11)

As a first approximation, the velocity in the surface layer is given by:

( )1 1 2*

d u hu

dt= (3.12)

and, assuming h1 is constant,

2*

11

u tuh

= (3.13)

Substituting Eq. 3.13 into Eq. 3.11 we have

2

12*

2g htu fB

′= (3.14)

This is the time it takes the interface to surface at the long side of the lake due to

the lateral tilting generated by the Coriolis force. The time it takes upwelling in

the upwind end of the lake is 4iT (Spigel and Imberger 1980), where

12LTc

= (3.15)

is the period of the fundamental mode and

12

1 2

1 2

h hc g'h h

⎛ ⎞= ⎜ +⎝ ⎠

⎟ (3.16)

is the phase speed of vertical mode one motions. The ratio of this two scales

Page 113: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 89

21

21

4 24 *

g ' h ctUT u fBL

= = = W S (3.17)

where 21

2′

=*

g hWu L

is the Wedderburn number and

2

=cS Bf

is the Burger number,

defines where upwelling happens first. If U>1, upwelling at the upwind end of the

basin occurs first; if U<1, upwelling at the left side of the wind (in the northern

hemisphere) occurs first.

Page 114: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

90 Chapter 3

Table 3.1. Characteristics of the internal bore generated after the full upwelling

event. hu and hd are the upstream and downstream depths of the interface, hu/hd is

the bore strength, Type is given by the regime defined by Rottman and Simpson

(1989), U and E are analytical estimates of speed of the bore and dissipation

generated by the bore, B is an estimate of buoyancy flux, Uobs is speed of the bore

estimated from observations, and BBobs is the buoyancy flux estimated from

observations.

Station hu (m)

hd (m)

u

a

hh

Type U (m s-1)

E (J m-1s-1)

B (J m-1s-1)

Uobs(m s-1)

Bobs(Jm-1s-1)

S5 40 0 ∞ Buoyant sur-face plume

0.26 179.9 44.9 0.22 51.2

S2 30 10 3 Breaking undular

0.18 13.3 3.3 0.22 -

S6 40 30 1.3 Laminar undular

0.15 0.4 0.1 - -

Page 115: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 91

Figure 3.1. Lake Constance (47°39’N, 9°18’E) bathymetry and location of Lake

Diagnostic Systems with thermistor chains and wind sensors. (Original map

courtesy of Martin Wessels).

Page 116: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

92 Chapter 3

0

3

6

9

12

W. s

peed

(m

s−1 ) a

0

90

180

270

360

Win

d di

r (d

eg)

0

10

20

30

40

50

b

Dep

th (

m)

0 2 4 6

x 10−3

290 295 300 305 310 315 320

0

10

20

30

40

50

c

Dep

th (

m)

Day in 2001 N (Hz)

Figure 3.2. 1-h averages of (a) wind speed (corrected to 10-m above the lake

surface) and direction (meteorological convention), (b) isotherm contours (contour

interval is 1.5°C and bottom contour is 6°C), and (c) buoyancy frequency at s7.

The dots at the left of (b) and (c) indicate the depth of the thermistors.

Page 117: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 93

0

3

6

9

12

W. s

peed

. (m

s−1 )

a

0

90

180

270

360

W d

ir. (

deg)

4.8255

4.826

4.8265

x 107

IPE

(kg

s−

2 ) b

10 20 30

290 295 300 305 310 315 320

2.1min

4.3min

8.5min

17.1min

34.1min

1.1h

2.3h

4.6h

Day in 2001

Peri

od (

min

/h)

c

event I event II

4−13 min

log2

Figure 3.3. (a) 10-min average of wind speed (solid line) and direction (dots) at

s2, (b) 10-s time series of integrated potential energy at s2, and (c) wavelet power

spectra of integrated potential energy at s2.

Page 118: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

94 Chapter 3

304.825 304.875

0

10

20

30

40

Day in 2001

Dep

th (

m)

a

300.025 300.075Day in 2001

b

Figure 3.4. Isotherm displacement showing high frequency internal waves (a)

during strong wind at s5 and (b) during the passage of the Kelvin wave trough at

s3.

Page 119: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 95

0

0.5

1a) s1

0

0.5

1b) s3

0

0.5

1c) s5

0

0.5

1d) s6

304 304.5 305 305.50

0.5

1e) s10

Day in 2001

Figure 3.5. Wind speed (ws, dotted line), integrated potential energy (IPE, dashed

line) and log2 of wavelet power of integrated potential energy in the 4- to 13-min

band (WIPE1, solid line) for some of the stations between days 304 and 305.5.

The variables were nondimensionalized as 16=ws ws ,

7

7 74.8252 10

4.8268 10 -4.8252 10− ×

=× ×

IPEIPE , and ( )( )

log2(WIPE1)- 27.1WIPE1=

41.2 -27.1. The shaded

area shows the period when high frequency waves were generated by shear

introduced by the wind.

Page 120: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

96 Chapter 3

0

0.5

1a) s1

0

0.5

1b) s3

0

0.5

1c) s5

0

0.5

1d) s6

299 299.5 300 300.50

0.5

1e) s10

Day in 2001

Figure 3.6. Wind speed (ws, dotted line), integrated potential energy (IPE, dashed

line) and log2 of wavelet power of integrated potential energy in the 4- to 13-min

band (WIPE1, solid line) for some of the stations between days 299 and 300.5.

The variables were nondimensionalized as 16=ws ws ,

7

7 74.8252 10

4.8268 10 -4.8252 10− ×

=× ×

IPEIPE , and ( )( )

log2(WIPE1)- 27.1WIPE1=

41.2 -27.1. The shaded

area shows the period when high frequency waves were generated by shear in the

trough of the basin-scale waves.

Page 121: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 97

20

25

30

Dep

th (

m)

a)

298 299 300 301 302

10

20

30

40

Dep

th (

m)

Day in 2001

b)

Figure 3.7. (a) Depth of the 9°C isotherm from field data (continuous line) and

ELCOM simulation (dashed line) at s3 located over the sill of Mainau. (b)

Regions in the time-depth plane with N larger than 0.002 Hz where the scaled Rig

(see text for definition) was below the critical value of 0.25.

Page 122: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

98 Chapter 3

306.35 306.4 306.45 306.5

5

10

15

20

25

30

35

40

Day in 2001

Dep

th (

m)

A B C

Figure 3.8. 0.5°C interval isotherm contours at s3 showing the fission of a

solitary-like wave. A thick contour shows the 9°C isotherm. A and B are vertical

mode 1 waves, C is a vertical mode 2 wave. The dots close to the vertical axes

indicate the vertical distribution of thermistors.

Page 123: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 99

0

0.5

1a) s5

0

0.5

1b) s3

0

0.5

1c) s2

0

0.5

1d) s1

305.5 306 306.5 307 307.5 308 308.50

0.5

1e) s6

Day in 2001

Figure 3.9. Wind speed (ws, dotted line), integrated potential energy (IPE, dashed

line) and log2 of wavelet power of integrated potential energy in the 13-min to

1.2-h band (WIPE2, solid line) for the stations in lake Überlingen and western end

of the main basin. The variables were nondimensionalized as 16=ws ws ,

7

7 74.8252 10

4.8268 10 -4.8252 10− ×

=× ×

IPEIPE , and ( )( )

log2(WIPE2)- 27.96WIPE2=

42.7 -27.96. The

arrows show the propagation of the solitary waves.

Page 124: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

100 Chapter 3

0

20

40

a

s3

Dep

th (

m)

306.3 306.50

3.5

7

log 2(W

13−

1.2)

(x10

11)

306.3 306.5

f

s3

307.6 307.8307.6 307.80

0.8

1.6

log 2(W

4−13

) (x

1011

)

b

s2

306.6 306.8306.6 306.8

e

s2

307.3 307.5Day in 2001

307.3 307.5

c

s1

306.8 307306.8 3070

0.8

1.6

log 2(W

4−13

) (x

1011

)

0

20

40

d

s1

Dep

th (

m)

307.1 307.30

3.5

7

log 2(W

13−

1.2)

(x10

11)

307.1 307.3

Figure 3.10. Isotherm displacements generated by the solitary waves (upper

panels) and the wavelet power in the 13-min to 1.2-h band (solid line, left axis)

and in the 4- to 13-min band (dashed line, right axis) (lower panels). (a), (b), and

(c) show the incident solitary waves from s3 to s1, and (d), (e), and (f) show the

reflected solitary waves from s1 to s3.

Page 125: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 101

0

0.5

1a) s5

304.5 305 305.5 306 306.50

0.5

1b) s7

Day in 2001

Figure 3.11. as Fig. 3.9 for stations s5 and s7.

Page 126: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

102 Chapter 3

Figu

re 3

.12.

Evo

lutio

n of

sur

face

wat

er te

mpe

ratu

re (

shad

ing)

and

vel

ocity

(th

in a

rrow

s) f

rom

ELC

OM

sim

ulat

ions

.

Thic

k ar

row

s sho

w th

e w

ind

spee

d.

Tim

e: 3

12.6

944

b

Tim

e: 3

13.7

222

d

Tim

e: 3

14

f

0.5

m s

−1

10 m

s−

1(°

C)

Tim

e: 3

12.1

944

a

Tim

e: 3

13.1

667

c

Tim

e: 3

13.8

611

e

68

1012

Page 127: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 103

313.7 313.8 313.9 314 314.1

0

10

20

30

40

50

60

70

80

90

100

Day in 2001

Dep

th (

m)

a) s5

314.3 314.4 314.5 314.6

0

10

20

30

40

50

Day in 2001

Dep

th (

m)

e) s5

314.1 314.2 314.3 314.4

0

10

20

30

40

50

b) s3

314.5 314.6 314.7 314.8

0

10

20

30

40

50

c) s2

314.8 314.9 315 315.1

0

10

20

30

40

50

Day in 2001

d) s1

Figure 3.13. 0.5°C interval isotherm contours showing the propagation of the

warm front through (a) s5, (b) s3, (c) s2, and (d) s1. e) shows the front reflected

from the sill as it passes s5. Shaded regions are magnified in Fig. 3.14. A thick

contour shows the isotherm 8.5°C in (a) 7.25°C in (b), and 8°C in (c), (d), and (e).

The dots close to the vertical axes indicate the vertical distribution of thermistors.

Page 128: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

104 Chapter 3

313.74 313.75 313.76 313.77

0

20

40

60

80

a) s5

313.8 313.85 313.9 313.95

0

20

40

60

b) s5

314.1 314.12 314.14

0

10

20

30

40

50

c) s3

Dep

th (

m)

314.48 314.5 314.52 314.54 314.56

0

20

40

60

d) s2

314.81 314.83 314.85 314.87

10

20

30

40

e) s2

Day in 2001314.76 314.77 314.78 314.79

0

10

20

30

40

50

f) s1

Day in 2001

Figure 3.14. Magnified views of shaded areas in Fig. 3.13. Isotherm contours

every 0.5°C with a thicker contour showing the isotherm 8.25°C in (a), 8.5°C en

(b), 7.25°C in (c), and 8°C in (d) (e), and (f). The dots close to the vertical axes

indicate the vertical distribution of thermistors. Note that the panels have different

scales.

Page 129: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

High-frequency motions in Lake Constance 105

313.9 313.95 314 314.05 314.1 314.15 314.2

10

15

20

25

30

35

40

45

50

55

Dep

th (

m)

a)

314.2 314.25 314.3 314.35 314.4 314.45

10

15

20

25

30

35

40

45

50

55

Day in 2001

Dep

th (

m)

b)

Figure 3.15. Isotherms displacement showing an undular internal bore as it passes

s6. (a) Incident bore, (b) reflected bore. The contour interval is 0.5°C and the thick

contour is 8.5°C. The dots close to the vertical axes indicate the vertical

distribution of thermistors.

Page 130: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

106 Chapter 3

Page 131: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

107

Chapter 4

Wind-shear Generated High-Frequency Internal Waves as

Precursors to Mixing in Lake Kinneret.

4.1 Abstract

Field data were used to estimate the wavelength, phase speed, direction of

propagation, frequency, and the vertical structure of high frequency internal

waves observed on the crests of basin-scale waves of Lake Kinneret during

periods of strong wind. Shear stability analysis indicates that these waves were

generated by shear in the surface mixing layer. The characteristics of the high

frequency internal waves changed within a wind event as the result of the

evolution of the background flow conditions following the deepening of the

surface layer and the propagation of the basin-scale internal waves. When the

background conditions were appropriate, the vertical structure of the unstable

mode was such that the perturbations generated visible sinuous internal waves that

in turn, modified the density profile in the metalimnion in such a way that

secondary shear instabilities were triggered. Poor coherence in temperature

records from stations 200 m apart suggested that the high frequency internal

waves were dissipated locally; these waves are thus a local mechanism allowing

energy to be drawn from the energized surface layer and transported to the

metalimnion where it sustains turbulence. Part of the energy extracted from the

surface layer was also returned to the mean flow in the metalimnion, high

frequency internal waves are therefore also a source of momentum for the

metalimnetic currents.

Page 132: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

108 Chapter 4

4.2 Introduction

Basin scale internal waves in stratified lakes have been widely studied

because they redistribute the momentum and energy introduced by the wind at the

surface of the lake (Mortimer 1974; Imberger 1998). High frequency internal

waves have also been detected in stratified lakes (Mortimer et al. 1968; Thorpe et

al. 1996) and are believed to contribute significantly to mixing (Imberger 1998).

Several generation mechanisms have been suggested but they can be grouped in

two distinct types. First, waves that evolve from hydraulic features such as basin

scale waves impinging on boundaries, basin scale wave shoaling over sloping

boundaries or basin scale waves simply steepening due to non linear effects

(Hunkins and Fliegel 1973; Wiegand and Carmack 1986; Thorpe et al. 1996).

These high frequency internal waves evolve into solitary waves and, as Boegman

et al. (2003) have shown, they can transport energy over large distances,

ultimately breaking when they shoal over a sloping bottom (Boegman et al.

2005b). Generally these waves contribute little to mixing within the metalimnion

in the interior of a lake. Second, there are sinusoidal waves that have their origin

in shear instabilities (Hamblin 1977; Stevens 1999). Boegman et al. (2003)

estimated that these waves dissipate over short distances by losing energy to the

background flow, but these authors were unable to conclusively connect the local

shear instabilities and the generated high frequency internal waves to the bursts of

turbulence observed in the metalimnion. Even though it is now recognized that the

bulk of the vertical mass flux in a lake is via the benthic boundary layer

(Goudsmit et al. 1997; Wüest et al. 2000; Saggio and Imberger 2001), it is still

important to understand the mixing in the metalimnion of the lake’s interior as

this not only transports mass, but also provides a rate of strain important for the

Page 133: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 109

bonding of chemical and biological particles. To evaluate the role of high

frequency internal waves in mixing in the interior of the lake, the generation

mechanisms, the propagation characteristics, and the dissipation mechanisms need

to be identified.

High frequency internal waves in Lake Kinneret were first described by

Antenucci and Imberger (2001b). They showed that the activity of these waves

was associated directly with the strength of the wind. The phase of the basin-scale

Kelvin internal wave had a modulating effect producing, for similar wind

conditions, stronger high frequency activity when the epilimnion was thinner.

These observations suggested that high frequency internal waves were generated

by shear in the surface layer induced by the wind. A linear stability analysis

disclosed that unstable modes with large growth rates had, in fact, similar

frequencies to those of the observed high frequency internal waves. Boegman et

al. (2003) defined more precisely the characteristics of the shear-unstable modes

associated with observations of high frequency internal waves in the crest of the

basin-scale internal Kelvin wave during periods of strong wind. Boegman et al.

(2003) also showed the existence of high frequency internal waves associated with

shear in the thermocline generated by high vertical mode basin-scale internal

waves, suggesting that the high frequency waves could also be generated by shear

instabilities at this level.

Shear instabilities have been widely investigated in order to explain the

excitation of gravity waves and other phenomena in the ocean and the atmosphere.

Most of the studies considered simplified two-dimensional profiles of background

velocity and density. Relatively simple background profiles such as constant shear

layer between constant velocity layers (Taylor 1931; Miles and Howard 1964;

Page 134: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

110 Chapter 4

Lawrence et al. 1991) and hyperbolic tangent function (Drazin 1958; Thorpe

1973; Davis and Peltier 1976) have been studied and have provided an

understanding of how the properties of shear unstable modes, such as wavelength,

frequency, phase speed, and growth rate, are related to the background flow

profiles. Solutions for these simplified profiles have also been used to suggest

shear instabilities as the generation mechanism for some waves observed in the

atmosphere (e.g., Lalas and Einaudi 1976) and in the ocean (e.g., Woods 1968;

Sutherland 1996). However, the onset of instability and the characteristics of the

unstable modes are sensitive to details in the background density and velocity

profiles (e.g., Howard and Maslowe 1973), but only few stability analyses have

been conducted using realistic background profiles in the atmosphere (De Baas

and Driedonks 1985), the ocean (Sun et al. 1998), and lakes (Boegman et al.

2003). In all these cases, however, a rigorous comparison of the characteristics of

the unstable modes and the internal waves has not been conducted due to limited

availability of data to estimate the characteristics of the observed waves.

In this paper, high spatial and temporal resolution field data are used to

estimate the frequency, wavelength, phase speed, direction of propagation, and

vertical modal structure of high frequency internal waves observed in the crest of

the internal Kelvin wave in Lake Kinneret. These estimates are then compared to

the results of a shear stability analysis of the background density and velocity

profiles. From this comparison we conclude that wind induced shear in the surface

layer is the generation mechanism for the high frequency internal waves. The

energy transfer supported by the high frequency internal waves is then

investigated in order to identify how these waves trigger mixing events in the

Page 135: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 111

metalimnion. The characteristics of the mixing events are described by Yeates et

al. (2007).

4.3 Field site

Lake Kinneret (Israel) is approximately 22 km long north-south and 15 km

wide west-east and had a maximum depth of about 39 m (Fig. 4.1) during the

summer of 2001, when the field experiment was conducted. The lake is

monomictic with a strong thermal summer stratification characterized by a

thermocline located about 16 m deep and a temperature difference of up to 8°C

across the metalimnion (Serruya 1975). Inflows, outflows, and density variations

due to salinity have a negligible effect on the dynamics at this time of the year. A

strong westerly breeze that reaches speeds of 15 m s-1 every afternoon excites a

set of basin-scale internal waves that are observed consistently during summer

(Serruya 1975; Antenucci et al. 2000; Gómez-Giraldo et al. 2006a). The latter

authors showed that the dominant natural mode is a vertical mode one, 22.6 h

period Kelvin wave that propagates anti clockwise around the entire lake. A

second dominant Poincaré wave mode, with a period close to 10.5 h, is also

observed and was shown by Gómez-Giraldo et al. (2006a) to be characterized by

two counter-rotating vertical mode one cells. These authors also showed that the

dominant periodicities in the wind mask the natural periods, shifting the observed

periods to 24 and 12 h respectively. Vertical mode 2 and 3 Poincaré waves with

periods close to 20 h have been detected in the northwestern region of the lake

(Antenucci et al. 2000).

Page 136: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

112 Chapter 4

4.4 Field equipment

An array of five closely spaced Lake Diagnostic Systems (LDSs) equipped

with thermistor chains, forming both a small and a large triangle, was deployed at

location B (Fig.1) to study the characteristics of the high frequency internal

waves. The distances between stations were planned to be 9 m for the small

triangle and 200 m for the larger triangle; the spacings were determined from

wavelength estimates obtained by Boegman et al. (2003). Their relative location is

presented in Fig. 4.1 and the actual distances and directions between the stations

are listed in Table 4.1. All the stations located at B had thermistors from a depth

of 0.75 m and were spaced every 0.75 m to the bottom. The thermistors had a

temperature accuracy of 0.01°C and a resolution of 0.001°C and were sampled

every 10 s.

Profiles of the background velocity and temperature microstructure were

measured with a Portable Flux Profiler (PFP; Imberger and Head 1994) equipped

with temperature sensors with a 0.001°C resolution and an orthogonal two-

component laser Doppler velocimeter with a 0.001 m s-1 resolution. With a

sampling frequency of 100 Hz and a fall velocity ~0.1 m s-1, the PFP sampled the

water column at scales as small as 1 mm. The majority of PFP profiles were

recorded near the centre of the large triangle (Fig. 4.1).

4.5 Analysis method: Linear stability analysis

To investigate whether the high frequency internal waves in the crests of

the internal Kelvin waves in Lake Kinneret originated from shear instabilities, as

proposed by Antenucci and Imberger (2001b) and Boegman et al. (2003), we used

the standard method to investigate the stability of sheared stratified media. A two-

Page 137: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 113

dimensional wave-like perturbation, characterized by a vertical velocity of the

form

( ) ( ) ( ){ }ˆw x,z,t w z exp ik x ct′ ⎡ ⎤= ℜ −⎣ ⎦ (4.1)

is introduced into a two-dimensional linear, incompressible, inviscid, stratified

shear flow with the Boussinesq and hydrostatic approximations, leading to the

Taylor-Goldstein equation:

( )

22

2 0zzzz

UNwU cU c

⎧ ⎫⎪ ˆk w⎪+ − − =⎨ −−⎪ ⎪⎩ ⎭⎬

i

i

(4.2)

where defines the vertical structure of the vertical velocity, k is the

horizontal wavenumber,

rˆ ˆ ˆw w iw= +

rc c ic= + is the complex phase speed, the horizontal

coordinate x has been orientated in the direction of k (without loss of generality), z

is the vertical coordinate defined positive upwards, t is time,

( ) ( )( )1 20N z g z= − ρ ∂ρ ∂ is the buoyancy frequency, ( )U z is the background

flow velocity, g is gravity, ( )zρ and 0ρ are the ambient and a reference densities,

and the subscripts indicate partial derivatives. For a known background density

and velocity profile and specified values of k, this equation was solved for and

c, and then the frequency

w

=r k crω and growth rate =i k ciω were calculated.

This model precludes three-dimensional primary instabilities that can grow from a

three-dimensional flow, but this type of instabilities are not expected in

geophysical flows because they are restricted to a very small region in the

parameter space of the background conditions (Smyth and Peltier 1990). Hence

we expect two-dimensional primary instabilities described by Eq. 4.1 to dominate.

Page 138: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

114 Chapter 4

We used the matrix eigenvalue method originally developed by Hogg et al.

(2001), and modified by Boegman et al. (2003), to solve Eq. 4.2 numerically for

velocity and density profiles obtained with the PFP. We specified zero vertical

velocity at the surface (the rigid lid condition) and at the bottom as boundary

conditions and searched for the complex phase speed, c, and the vertical velocity

structure, , of unstable modes with wavelengths w 2= kλ π at 2-m intervals

between 2 and 60 m. Although the code can solve a generalized form of the

Taylor-Goldstein equation that includes viscosity and diffusivity, the turbulence in

the metalimnion of Lake Kinneret is triggered by the events we were seeking;

their precursor was a laminar water column (Saggio and Imberger 2001) so

viscosity and diffusion could be neglected.

To analyse the two-dimensional instabilities, the horizontal velocity

profiles were projected onto 32 vertical planes with directions separated by 11.25°

and Eq. 4.2 was solved for each plane. The directional nature of the instabilities

was then recovered by combining the solutions in wavenumber space.

4.6 Observed background conditions

The periodic westerly afternoon breeze acting over lake Kinneret (Fig.

4.2a) excited an exceptionally regular pattern of basin-scale internal waves

(Gómez-Giraldo et al. 2006a) (Fig. 4.2b). The largest component in the observed

oscillations was the result of the near resonant interaction of the 22.6 h natural

period Kelvin wave with the 24 h dominant periodicity in the wind forcing. At

location B, the crest of the Kelvin wave was in phase with the wind so the

thickness of the epilimnion always decreased when the wind was strongest. This

produced a strong shear in the surface layer, pointing to shear instability as the

Page 139: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 115

generation mechanism for the high frequency internal waves riding on the crests

of the Kelvin wave (Fig. 4.2b) (Antenucci and Imberger 2001b). Figures 4.2c,d

show magnified views of two high frequency events in the afternoons of days 179

and 182. The high frequency internal waves had amplitudes up to 1 m, were

vertically coherent and occurred in packets. Power spectra (Bendat and Piersol

1986) of the isotherm displacements, measured with all LDS’s, showed that the

dominant Eulerian frequency in the high frequency range was about 0.0056 Hz

(180 s period), as illustrated in Fig. 4.3 for the 23°C isotherm at TB1. The power

spectra of the 23°C isotherm vertical displacements at the other LDS stations were

identical and the same frequency also dominated the high frequency oscillations

of other isotherms throughout the water column.

Miles (1961) and Howard (1961) defined a quantitative criteria for

instability showing that small perturbations may grow, extracting energy from the

background flow, only if the gradient Richardson number

( )22 0 25gRi N z .= ∂ ∂ <U somewhere in the water column. Figure 4.4 shows the

regions of the water column where Rig falls below the critical value of 0.25 for the

afternoon of day 179, when the high frequency internal wave activity was strong

and the background conditions were continuously monitored with PFP casts. In

the surface layer Rig dropped below the critical value (Fig. 4.4a) due to the east-

directed shear introduced by the westerly wind (Fig. 4.4b) and because of the

weak stratification. Rig also decreased below 0.25 in the strongly stratified

metalimnion due to the strong shear associated with the north-south currents (Fig.

4.4b) generated by the high vertical mode basin-scale Poincaré waves described

by Antenucci et al. (2000). One of the objectives of this analysis is to determine

Page 140: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

116 Chapter 4

the region primarily responsible for the generation of the high frequency internal

waves.

4.7 Characteristics of the observed high frequency waves

Antenucci and Imberger (2001b) noticed that the high frequency waves

appeared only sporadically. The 23°C isotherm vertical displacements series from

all 5 stations was band pass filtered with a narrow filter around a frequency of

0.0056 Hz to identify in what periods the high frequency internal waves were

present. Figure 4.5a shows that the waves were generated at all 5 stations during

the same periods, mainly when the wind was strong. This suggests that the

generation mechanism had a spatial scale larger than 200 m. However, Figs.

4.5b,c show that the wave packets were only similar for the stations in the small

triangle, but they were less similar for the stations that formed the large triangle.

This is a first indication that the waves were generated and dissipated locally over

distances of the order of 100 m.

The strongest high frequency internal wave activity in the metalimnion

was not concurrent with the maximum wind speed. Further, there was a delay of

about 4 h between the start of the wind and the appearance of the high frequency

internal waves in the metalimnion. This indicates that the background flow

conditions, which were modulated by the passage of the basin-scale internal

waves, had a strong effect on the generation of high frequency internal waves. It

was also likely that the characteristics of the high frequency internal waves

changed within a wind event following changes in the wind and in the background

flow, so we investigated the characteristics of these waves during three periods

within the wind event observed in the afternoon of day 179. The start and end

Page 141: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 117

times of the periods and the number of casts during each period are presented in

(Table 4.2). Period 1500 was characterized by strong wind, very weak high

frequency internal waves activity in the metalimnion and some deepening of the

mixed layer. The maximum wind speed occurred during period 1630, which

included the first PFP casts after the beginning of the high frequency internal

wave activity. Period 1900 covered the maximum elevation of the metalimnion

and strong high frequency internal wave activity.

Spectral analysis of the 23°C isotherm displacements, shown in Fig. 4.6,

indicates that there was not a clear peak for period 1500, when the internal wave

activity in the metalimnion was very weak. Power spectral density of the 27°C

isotherm (also in Fig. 4.6) indicates that the more active surface mixing layer also

lacked a dominant frequency for this period. During periods 1630 and 1900, the

displacements in the metalimnion were dominated by oscillations with frequencies

of 0.0058 and 0.0042 Hz (172 and 238 s) respectively.

The propagation characteristics of the high frequency internal waves were

estimated from a coherence and phase analysis (Bendat and Piersol 1986) of the

23°C isotherm displacement for the three stations in the small triangle. The

results, presented in (Table 4.3) for periods1630 and 1900 (the lack of a confident

peak prevents calculation for the period 1500), show that the high frequency

internal waves properties changed within one wind event. The coherence between

the signals from the stations in the large triangle was small, invalidating any

results from the phase analysis applied to the large triangle and supporting the

conclusions of Boegman et al. (2003), that the high frequency internal waves were

generated and dissipated locally.

Page 142: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

118 Chapter 4

The vertical structure of the high frequency internal waves was extracted

from the vertical velocity profiles measured with the PFP. Figure 4.7 shows

representative profiles for each one of the three periods considered. It was not

possible to isolate the motion due to the dominant high frequency internal waves

summarized in (Table 4.3), but we expect the vertical velocity profiles to be

strongly correlated to these waves, as they dominated the isotherm response. The

profile for period 1500 exhibited strong small-scale fluctuations above 5 m, i.e., in

the surface mixing layer. There was a local maximum at 7 m depth and then the

vertical velocity decreased rapidly below this level, being almost zero in the

metalimnion. During the period 1630, there were also strong small-scale

fluctuations above 5 m and the vertical velocity again decreased with depth, but

there was a region with considerable vertical velocity between 13 and 20 m deep.

The vertical velocity profile for period 1900 exhibited, in addition to the

fluctuations in the top 5 meters, an intermediate minimum at 15 m depth and two

local maxima above and below this level; the manifestation of the changes in the

vertical velocity distribution on the isotherm displacement may be noted in Fig.

4.4b.

4.8 Results of the stability analysis The background shear and density profiles required as input to Eq. 4.2

were obtained, for each one of the three periods, by calculating mean isotherm

depths and the corresponding isopycnal velocities from the PFP casts. In this way,

the fluctuations in the individual profiles generated by the high frequency internal

waves, were smoothed out. The density and velocity in the top 3 or 4 meters,

where PFP data was unavailable, were estimated by extrapolating PFP data to the

Page 143: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 119

surface. Figure 4.8 presents the density and velocity profiles, together with the

associated gradient Richardson number (Rig) and direction of the shear. The

evolution of the density profile in Figure 4.8a shows a deepening of the diurnal

thermocline between periods 1500 and 1630, as a result of the wind-induced

mixing. Between periods 1630 and 1900, the entire stratification was lifted by the

basin-scale internal waves. Fig. 4.8b reveals that, in addition, the metalimnetic jet

became stronger and shifted toward the surface. There were two depths where the

associated Rig decreased below 0.25; as the afternoon progressed, these regions

became larger and moved closer together (Fig. 4.8d,e,f). The shallow region, in

the surface layer, was produced by the gradient in the zonal component of the

velocity due to the shear stress introduced by the wind at the surface of the lake.

Consequently, the shear was directed to the East (Fig. 4.8d,e,f). The second,

deeper region, in the metalimnion, was produced by the meridional shear in the

metalimnion. This shear was directed to the north at the level where Rig was

minimum, but varied between the northwest and the northeast over the small Rig

region (Fig. 4.8d,e,f). We expect shear instabilities to have been generated in

those two depth ranges and, accordingly, unstable modes should have propagated

predominantly along those two dominant directions. In principle, the mode with

the fastest growth rate should be the one responsible for the observed high

frequency internal waves.

The averaged profiles were then interpolated to the nodes of the numerical

grid using cubic splines and the Taylor-Goldstein equation was solved

numerically as explained earlier. Ten, 5, and 2.5 cm grid sizes were used for each

period in order to investigate the sensitivity of the numerical code to the

discretization. As an example, Fig. 4.9a shows the growth rate of unstable modes

Page 144: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

120 Chapter 4

for direction 101.25° during the period 1500, as calculated with the 5 cm grid, and

illustrates how the modes were separated into four types. Modes A (Fig. 4.9a)

were not sensitive to the grid size, their growth rate had a well-defined peak in the

range of wavenumbers considered, and their critical level (where U equals cr) was

located in the surface mixing layer. Modes B did not have a defined growth rate

peak in the range of wavenumbers considered, were sensitive to the grid size in

some cases, and their critical level was located in the metalimnion. Modes C were

isolated, i.e., did not exist for neighbouring wavenumbers, and were very sensitive

to the grid size. Modes with small growth rates were classified as D and were not

further considered. Modes C were most likely a numerical version of the modes

produced in the interface of homogenous layers, which were first studied by

Taylor (1931) and called T-modes by Caulfield (1994); these modes were the

results of the numerical discretization. Numerical investigations directed to

evaluating the code behaviour for simple shear and density profiles (Andy Hogg,

pers. comm.) confirmed this limitation of the numerical scheme. The character of

the B modes was unclear; some could be numerical T-modes, as suggested by

their sensitivity to the grid size; others, insensitive to the grid size, could have

been modes associated with thin regions of elevated shear so that the wavelength

at which they reach the maximum growth rate was smaller than the shortest

wavelength analyzed (2 m) and no peak in growth rate was observed in the range

of wavelengths considered. The thin shear regions that generated these modes

were probably generated by fluctuations in the background profiles that were not

removed when smoothing the background conditions. These modes were an

artificial consequence of the numerical approximation to the real background

profiles and should be disregarded. A third possibility is that these modes were

Page 145: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 121

real; however, they could not be associated to the high frequency waves because

their wavelength (~ 2 m) and frequency (~ 1×10-2 Hz) differed by a factor of

about 10 from those of the observed high frequency internal waves. In what

follows, we study further the characteristics of the A modes and compare them to

those of the observed high frequency waves.

Figure 4.10 shows that, during the three periods considered, the faster

growing shear-unstable perturbations were orientated along a direction between

90 and 135°, in good agreement with the estimated direction of propagation of the

high frequency internal waves (Table 4.3). The faster growing modes for period

1500 orientated along the direction 101.25°. The modes that most likely

dominated are identified in Fig. 4.9a; their vertical structure is presented in Fig.

4.9b-e and their propagation characteristics are presented in (Table 4.4).

Unfortunately, the propagation characteristics could not be inferred from the

observed high frequency internal waves as the metalimnetic isotherm

displacements were small, but the vertical velocity profile agreed well with the

PFP measurement (Fig. 4.7a): both exhibited a local maximum of vertical velocity

at about 7 m deep and a rapid decay below that level. The critical level of these

modes was located around 4.7 m deep and the corresponding local minimum in

amplitude and shift in phase associated with the critical level (De Baas and

Driedonks 1985) were observed.

For period 1630, the faster growing A modes were orientated along the

112.5° direction. The four modes with the highest growth rates are identified in

Fig. 4.11a; their vertical structure is presented in Fig. 4.11b-e, and their

propagation characteristics in (Table 4.4). Both the propagation characteristics

(Tables 4.3 and 4.4) and vertical structure (Figs. 4.11b-d and 4.7b) agreed with the

Page 146: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

122 Chapter 4

observations. The fastest growing modes for period 1900 orientated along the

101.25° direction and are presented in Fig. 4.12. Modes I and II have a slightly

larger growth rate, but the vertical velocities associated with them were very small

below 10 meters. Modes III and IV have slightly smaller growth rate but they

support vertical velocity perturbations to a depth of approximately 23 m. This,

together with a closer match with the propagation characteristics of the observed

high frequency internal waves links mode IV with the fluctuations observed

during this period. In addition, any reduction in the growth rate due to the viscous

effects should be smaller in this mode than in modes I, II, and III due to its

smaller frequency and larger wavelength.

4.9 Perturbation kinetic energy

We examined the flux of kinetic energy of the perturbations to illustrate

how these unstable modes participated in the energy flux path. Considering that

dissipation is negligible at the scales of the perturbations and that the vertical

transport of kinetic energy is dominated by the work done by pressure, the

perturbation kinetic energy (P.K.E.) equation becomes (Sun et al. 1998; Kundu

and Cohen 2002)

( )0 0

1′ ′ ′ ′ ′ ′= − − −zt zgP.K .E u w U w p wρρ ρ

(4.3)

where the angle brackets denote averages over one horizontal wavelength. The

terms on the right hand side are, from first to last, shear production, work done by

gravity, and convergence of the vertical energy flux. The terms u , , and p′ ′ ′ρ are

the horizontal velocity, density, and pressure perturbations due to the unstable

mode and are given by

Page 147: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 123

[ ]( ) ( ) ( ) ( ) ( ){ }′ ′ ′ ⎡ ⎤ ⎡ ⎤= ℜ −⎣ ⎦ ⎣ ⎦ˆˆ ˆu , , p x,z,t u z , z , p z exp ik x ctρ ρ (4.4)

with

= ziˆ ˆu wk

(4.5)

0ˆ ˆi N wg

2ρρ =σ

(4.6)

0 2zi iˆ ˆp U wk k

σρ zw⎧ ⎫= +⎨ ⎬⎩ ⎭

(4.7)

and U kσ ω= − is the Doppler shifted complex frequency.

Figure 4.13 shows the terms on the right hand side of Eq. 4.3 for mode IV

during the period 1900, with from Fig. 4.12e, and c and k from Table 4.4.

Clearly, kinetic energy is drained from the mean velocity field by shear

production at the critical level, which is located in the surface mixing layer (Fig.

4.13b). Some of this energy was lost to potential energy (Fig. 4.13c) and some

was transported away from this level (Fig. 4.13d); some P.K.E. was transported to

the region between 12 and 14 m (Fig. 4.13d) were it was transferred back to the

mean flow (Fig. 4.13b).

w

4.10 Discussion

Close agreement between the characteristics of the observed high

frequency internal waves and the shear-unstable modes confirmed that such waves

were generated by shear instabilities. Despite the presence of a region with Rig

below 0.25 in the metalimnion, the modes responsible for the high frequency

internal waves were unstable to shear in the surface layer due to the wind-

generated stress as initially suggested by Antenucci and Imberger (2001b). The

Page 148: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

124 Chapter 4

initial numerical results of the linear and inviscid Taylor-Goldstein equation

included some B modes with growth rates higher than those of the modes A

associated with the high frequency internal waves (Fig. 4.9). Although they were

candidates as a source of high frequency internal waves, there are several reasons

why these modes did not dominate in the field observations over the A modes.

Some were a numerical version of the T-modes (Caulfield 1994) or artificially

generated by the limitations in our approximation to the real background

conditions as discussed above. In the case these modes were real solutions to Eq.

4.2 and assuming that the background profiles were properly smoothed, their

small wavelength (~ 2 m) and high frequency (~ 1×10-2 Hz) were too different

from those observed to be realistic candidates. The absence of visible oscillations

related to the B modes can be explained by two different mechanisms. First,

viscous effects may become important for these modes due to their small

wavelengths and high frequencies and reduce their growth rates. Boegman et al.

(2003) showed that the inclusion a constant eddy viscosity in the generalized

Taylor-Goldstein equation does not lead to good agreement with observations.

However, these authors estimated that the dissipation rate overcomes the growth

rate for modes with small wavelength and high frequency. Second, nonlinear

numerical simulations carried out by Sutherland and Peltier (1994) revealed that

the vortices formed by the growth of unstable modes in a region of large N are

rapidly strained leading to a cascade of vorticity to smaller scales. A turbulent

patch centred at 14.5 m depth, which is at the critical level of the B modes

predicted for period 1900, is reported by Yeates et al. (2007). This suggests that,

despite not being related to the observed high frequency internal waves, some of

the B modes were real. However, the turbulent properties of the patch (Yeates et

Page 149: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 125

al. 2007) indicate that the turbulent events originated by the collapse of these

modes were not very intense and did not contribute significantly to the buoyancy

flux.

We showed how the characteristics of the unstable modes changed during

a typical wind event. Of particular interest is the evolution of local maxima of the

vertical velocity in the metalimnion, away from their critical level. The square

root of the term inside the brackets in Eq. 4.2 can be interpreted as the complex

vertical wavenumber of the unstable mode and this controls the vertical shape of

the mode; its real component gives the decay rate of the amplitude of the mode

with distance from the critical level and its imaginary component gives the

waviness of the mode. Although it is difficult to isolate their individual

contributions for profiles like those observed in Lake Kinneret (Fig. 3.8),

variations in N, (U-cr), and 2d U dz2 influence the vertical structure of a given

mode. For this reason, the use of simple analytical background profiles with

constant density and velocity profiles away from the shear layer, suitable to

understand basic aspects of shear unstable modes (e.g., Hazel 1972; Davis and

Peltier 1976) and to explain radiation of internal waves away from the generation

zone in the ocean (Sutherland 1996) and the atmosphere (Sutherland and Peltier

1995), provides limited information about the vertical structure of the mode.

Therefore, the simplified profiles are inadequate to predict the direct effects of the

unstable mode far from the generation zone, like the presence of high frequency

internal waves in the metalimnion of Lake Kinneret. We have shown that, despite

modes being associated with shear in the surface mixing layer, the high frequency

internal waves activity in the metalimnion was not observed from the start of the

strong wind event, but required the evolution of the background profiles through

Page 150: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

126 Chapter 4

the afternoon. Antenucci and Imberger (2001b) noted that the amplitude of the

basin scale internal waves and their phase, relative to the wind, modulated the

generation of high frequency waves and he suggested that this was a consequence

of the modification of the thickness of the surface layer and hence of the mean

shear. We can add now that it was also due to the changes in the entire

background flow including the closer displacement of the metalimnion to the

critical level of the unstable modes and the evolution of the metalimnetic jet.

The vertical velocity profiles in Fig. 4.7 exhibit some small-scale

fluctuations that are not directly related to the vertical structure of the unstable

modes. Most of them were located in the top 5 meters and can be associated to the

stirring generated by the wind and the billowing generated by unstable modes

with vertical structures confined to a thin vertical region near the surface. The

fluctuations near the bottom can be associated with turbulence in the benthic

boundary layer. Of particular interest are the fluctuations that the profile in (Fig.

4.7c) exhibits just below15 m depth. They could not be generated by shear

unstable modes, even those previously disregarded, because none of the modes

had a critical level at this depth. In addition, the region in the metalimnion with

minimum Rig values, where the B modes gained their energy from, was located

above 15 m. This suggests yet a different generation mechanism for these

particular velocity fluctuations, observed just below 15 m.

Figure 4.14a shows the vertical velocity perturbation due to the unstable

mode in Fig. 4.12e after it has grown from an initial (arbitrary) maximum

amplitude of 0.02 m s-1 at 4 m depth during 1000 s until it reached a maximum

amplitude of 0.05 m s-1, which is close to the velocity measured by the PFP (Fig.

4.7c) at that level. Figure 4.14b shows the associated density perturbation

Page 151: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 127

generated by this mode at this stage of growth (calculated with Eqs. 4.4 and 4.6),

and Fig. 4.14c shows the total density obtained superimposing the density

perturbation in Fig. 4.14b on the background density in Fig. 4.8a. This last panel

shows that the density profile is just about to become statically unstable between

14.8 and 15 m depth. Beyond this time, convective instabilities can develop

creating the small-scale fluctuations observed in Fig. 4.7c. However, Yeates et al.

(2007) showed that the reduction in the density gradient at this depth reduces Rig

to a level where the flow becomes shear unstable before the convective

instabilities develop. These authors also show that the turbulent motion resulting

from the collapse of these secondary unstable modes was very active and

produced large buoyancy fluxes.

Unlike the statically unstable profiles observed in the surface mixing layer,

which provide evidence of the overturning of the large billows generated by the

primary unstable modes, LDS data did not reveal any statically unstable profile in

the metalimnion that could have exposed the existence of the secondary

instabilities. This is because the characteristic length scale of the overturns at that

level was approximately 0.05 m (Yeates et al. 2007), far smaller than the 0.75 m

separating two consecutive thermistors. Also, the size of the stationary bin used

for the estimation of the turbulent properties in the water column, which indicates

the size of the turbulent patches, was only about 0.40 m.

Figure 4.13c shows that there is no net change of perturbation potential

energy around 15 m, but this does not contradict the suggested generation of a

secondary instability. This instability is triggered by the oscillations of the

density field, while the calculation of the perturbation energy flux requires

integration over a wavelength. The energy flux described by Fig. 4.13 can affect,

Page 152: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

128 Chapter 4

however, the mean flow by feeding the strong currents observed in the

metalimnion. In fact, as pointed out by Yeates et al. (2007), the currents generated

by the three gravest vertical basin-scale modes were not sufficient to explain the

strong jet in the metalimnion. A similar mechanism was proposed by Sutherland

(1996) for the generation of the deep equatorial undercurrents, with the difference

that in that case the momentum is transported by free high frequency internal

waves.

The horizontal transport of energy by the high frequency internal waves is

limited to a distance of the order of hundreds of meters (Boegman et al. 2003).

They were observed at all the stations considered in this study because the

structure of the basin-scale waves, the bathymetry and the strength of the wind

vary over scales larger than the dimensions of the LDSs array. However, the loss

of coherence over a distance of 200 m (the separation of the thermistor chains in

the large triangle) indicates that the high frequency internal waves were generated

and dissipated locally. The dissipation length scales, estimated as =D r iL c ω ,

were 81 and 67 m for modes IV (those most likely responsible for the high

frequency internal waves) during periods 1630 and 1900 respectively.

4.11 Conclusions

It was shown that in a stratified lake energy is extracted from the mean

flow by shear instabilities in the surface layer, then transported by high frequency

internal waves to the metalimnion where the straining due to the combination of

these high frequency internal waves and the mean basin scale flow may lead, as

shown by Yeates et al. (2007), to elevated levels of dissipation through the

excitation of secondary instabilities. In addition, part of the energy extracted from

Page 153: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 129

the surface layer is returned to the mean flow in the metalimnion providing a

momentum input there. In this way the shear generated high frequency internal

waves are responsible for a vertical transport of energy from the surface layer to

the metalimnion and thus contribute significantly to the erosion of the

stratification and to the horizontal advection. This has important consequences for

the energy flux in stratified lakes, but a quantification of such energy fluxes is still

pending as well as a parameterization of these processes in coarse grid numerical

models.

4.12 Acknowledgments

The field experiment was funded by the Israeli Water Commission, via the

Kinneret modeling Project, and was conducted by the Field Operations Group at

the Centre for Water Research (CWR) and the Kinneret Limnological Laboratory

(KLL). The authors acknowledge Leon Boegman and Andy Hogg for the

numerical code used in the linear stability analysis. We also thank Peter Rhines,

Peter Yeates, Andy Hogg, and Leon Boegman for very constructive discussions

and comments on an earlier version of this manuscript. The first author

acknowledges the support of an International Postgraduate Research Scholarship

and an ad-hoc CWR scholarship. This paper represents Centre for Water Research

reference ED 1672 AG.

Page 154: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

130 Chapter 4

Table 4.1. Characteristics of the configuration of the LDS array.

Line Distance (m) Direction (°)

TB1-TB2 9.1 211

TB1-TB3 8.1 144

TB2-TB3 9.6 83

TB1-TB4 202 212

TB1-TB5 195 143

TB4-TB5 225 86

Page 155: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 131

Table 4.2. Periods of analysis of the characteristics of the high frequency internal

waves.

Period Time of first PFP

cast

Time of last PFP

cast

Number of PFP

casts

1500 14:39 15:40 12

1630 16:08 17:05 11

1900 18:39 19:35 11

Page 156: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

132 Chapter 4

Table 4.3. Characteristics of measured waves. θ is the direction of propagation.

Coh2 ϕ (rad)

Period rω

(Hz) TB1-

TB2

TB1-

TB3

TB2-

TB3

TB1-

TB2

TB1-

TB3

TB2-

TB3

λ

(m)

cr

(m s-1)

θ

(°)

1630 0.0058 0.83 0.87 0.65 -0.024 1.844 2.014 24 0.14 120

1900 0.0042 0.94 0.80 0.74 0.101 2.537 2.316 19 0.08 124

Page 157: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Chapter 4 133

Period 1500 Period 1630 Period 1900

Mode λ (m) rc (m s-1) rω (Hz) iω (rad s-1) λ (m) rc (m s-1) rω (Hz) iω (rad s-1) λ (m) rc (m s-1) rω (Hz) iω (rad s-1)

I 20 0.16 0.0079 0.0037 18 0.13 0.0073 0.0016 12 0.08 0.0068 0.0014

II 22 0.16 0.0073 0.0036 20 0.13 0.0066 0.0016 14 0.08 0.0059 0.0014

III 24 0.16 0.0068 0.0033 22 0.13 0.0060 0.0016 16 0.08 0.0051 0.0013

IV 26 0.16 0.0064 0.0030 24 0.13 0.0055 0.0016 18 0.08 0.0046 0.0012

Table 4.4. Propagation characteristics of the fastest growing modes of the three periods considered. = 2π kλ is the wavelength.

Page 158: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

134 Chapter 4

0 5 10 150

5

10

15

20

Easting (km)

Nor

thin

g (k

m)

10

20

30

B

TB1

TB2 TB3

TB4TB5

Figure 4.1. Map of Lake Kinneret. The filled circle at B indicates the location of

the 5-LDS array. The inset sketches the relative location of the LDSs (filled

circles), with distances and direction of the lines specified in Table 4.1. The cross

near the centre of the large triangle indicates the approximate location of the PFP

casts. The origin of the map coordinate system is situated at 35.51°N, 32.70°E.

Page 159: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 135

0

4

8

12

16W

ind

sp. (

m s

−1 )

N

E

S

W

N

Win

d di

r. (

°)a

179 179.5 180 180.5 181 181.5 182 182.5 183

5

10

15

20

25

Day in 2001

Dep

th (

m)

c db

179.79 179.82

6

8

10

12

14

16

18

Day in 2001

Dep

th (

m)

c

182.68 182.71Day in 2001

d

Figure 4.2. Wind and temperature observations at TB1. (a) Ten-minute average

wind speed and direction. (b) Temperature contours at 2°C intervals; the bottom

isotherm is 17°C. (c) and (d) Magnified views of shaded regions c and d in (b)

with temperature contours at 1°C intervals; the bottom isotherm in both panels is

18°C.

Page 160: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

136 Chapter 4

10−6

10−5

10−4

10−3

10−2

10−1

10−4

10−2

100

102

104

106

Frequency (Hz)

Spec

tral

den

sity

(m

2 Hz−

1 )

24 h 12 h

180 s 100 s

Figure 4.3. Power spectral density of 23°C isotherm vertical displacements at

station TB1 for the entire length of the record. The vertical dotted lines indicate

the periods of the dominant basin-scale (24 and 12 h) and high frequency waves

(180 s), the vertical dashed line indicates the maximum buoyancy frequency, and

the dashed lines near the bottom define confidence at the 95% level.

Page 161: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 137

Dep

th (

m)

a

0

5

10

15

20

Day in 2001

Dep

th (

m)

b

179.6 179.65 179.7 179.75 179.8 179.85 179.9

0

5

10

15

20

0 45 90 135 180 225 270 315 360

Figure 4.4. (a) Regions with Richardson number smaller and larger than 0.25 are

coloured in black and grey respectively. The white lines are isotherms at 2°C

intervals. (b) Direction of the dominant shear (following the oceanographic

convention). The white lines are isotherms at 2°C intervals. Small vertical lines at

the top of each panel indicate the times at which the PFP casts were made, from

which the Richardson number and shear were determined.

Page 162: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

138 Chapter 4

170 172 174 176 178 180 182

0

2

4

6

8

TB1

TB2

TB3

TB4

TB5

Day in 2001

Ver

tical

dis

plac

emen

t (m

) a

170.65 170.7 170.75

0

2

4

6

8

TB1

TB2

TB3

TB4

TB5

Day in 2001

Ver

tical

dis

p. (

m)

b

179.81 179.86 179.91

TB1

TB2

TB3

TB4

TB5

Day in 2001

c

Figure 4.5. (a) 23°C isotherm vertical displacements band-pass filtered around

0.0056 Hz for the entire record period. (b) and (c) Magnification for two periods

of strong high frequency internal wave activity.

Page 163: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 139

10−4

10−3

10−2

10−1

10−4

10−3

10−2

10−1

100

101

102

103

Frequency (Hz)

Spec

tral

den

sity

(m

2 Hz−

1 )

238 s 172 s

23°C 150023°C 163023°C 190027°C 1500

Figure 4.6. Power spectra of the 23°C isotherm displacements for periods 1500,

1630, and 1900, and of the 27°C isotherm displacement for period 1500. Spectra

have been smoothed in the frequency domain to increase confidence. The dashed

lines near the bottom define confidence at the 95% level.

Page 164: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

140 Chapter 4

−0.05 0 0.05

0

5

10

15

20

25

30

a

Dep

th (

m)

−0.05 0 0.05

b

Vertical velocity (m s−1)−0.05 0 0.05

c

Figure 4.7. Selected vertical velocity from PFP data for period (a) 1500, (b) 1630,

and (c) 1900. Shaded areas indicate the extent of the metalimnion.

Page 165: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 141

996 998 1000

0

5

10

15

20

25

30

ρ (kg m−3)

Dep

th (

m)

a

−0.2 0 0.2 U

0 (m s−1)

b

−0.2 0 0.2 U

90 (m s−1)

c d e

150016301900

f

0 45 90 135 180

Figure 4.8. Mean background profiles for the three periods were stability was

investigated during day 179. (a) Density, (b) and (c) meridional (U0) and zonal

(U90) components of velocity. (d) Shading indicates the direction of shear for

period 1500 at those depths where Rig < 0.25. The direction was calculated

as 1 0 9tan dU dUdz dz

θ − ⎛ ⎞⎛ ⎞ ⎛ ⎞= ⎜ ⎟ ⎜ ⎟⎜⎝ ⎠ ⎝ ⎠⎝ ⎠

0⎟ . (e) and (f) Same as (d) for periods 1630 and

1900 respectively.

Page 166: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

142 Chapter 4

0 0.2 0.4 0.6 0.8 1 1.2 1.40

1

2

3

4

5

6x 10

−3

k (rad m−1)

ωi (

rad

s−1 )

a) Growth rate

III

IIIIV

ABCD

−1 0 1

0

5

10

15

20

25

30

Amplitude

Phase/π

b) I

Dep

th (

m)

−1 0 1Amplitude

Phase/π

c) II

−1 0 1Amplitude

Phase/π

d) III

−1 0 1Amplitude

Phase/π

e) IV

Figure 4.9. (a) Growth rate of unstable modes for direction 101.25° during period

1500. Modes A have a well-defined peak in the growth rate in the range of

wavenumbers considered and are insensitive to the grid size in the numerical

solution of the Taylor-Goldstein equation. Modes B do not have a well-defined

peak in the growth rate in the range of wavenumbers considered and are

sometimes sensitive to the grid size. Modes C are isolated and sensitive to the grid

size. Modes D have small growth rates. (b), (c), (d), and (e) Vertical structure of

the modes I, II, III, and IV in terms of relative amplitude (solid line) and phase

(dashed line).

Page 167: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 143

k

l

a

−0.5 0 0.5

−0.5

0

0.5

0.0000 0.0020 0.0040

ωi (rad s−1)

k

b

−0.5 0 0.5

0.0000 0.0020

ωi (rad s−1)

k

c

−0.5 0 0.5

0.0000 0.0016

ωi (rad s−1)

Figure 4.10. Maximum growth rate of unstable modes in horizontal wavenumber

space for (a) period 1500, (b) period 1630, and (c) period 1900. The panels are

limited to the region –1 < k, l <1, where all the fastest growing modes are located.

Page 168: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

144 Chapter 4

0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.50.4

0.6

0.8

1

1.2

1.4

1.6

1.8x 10

−3

k (rad m−1)

ωi (

rad

s−1 )

a) Growth rate III

IIIIV

−1 0 1

0

5

10

15

20

25

30

Amplitude

Phase/π

b) I

Dep

th (

m)

−1 0 1Amplitude

Phase/π

c) II

−1 0 1Amplitude

Phase/π

d) III

−1 0 1Amplitude

Phase/π

e) IV

Figure 4.11. (a) Growth rate of unstable modes type A for direction 112.5° during

period 1630. (b), (c), (d), and (e) Vertical structure of the modes I, II, III, and IV

in terms of relative amplitude (solid line) and phase (dashed line).

Page 169: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 145

0 0.2 0.4 0.6 0.8 1 1.2 1.42

4

6

8

10

12

14

16x 10

−4

k (rad m−1)

ωi (

rad

s−1 )

a III

IIIIV

−1 0 1

0

5

10

15

20

25

30

Amplitude

Phase/π

b) I

Dep

th (

m)

−1 0 1Amplitude

Phase/π

c) II

−1 0 1Amplitude

Phase/π

d) III

−1 0 1Amplitude

Phase/π

e) IV

Figure 4.12. (a) Growth rate of unstable modes type A for direction 101.25°

during period 1900. (b), (c), (d), and (e) Vertical structure of the modes I, II, III,

and IV in terms of relative amplitude (solid line) and phase (dashed line).

Page 170: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

146 Chapter 4

−1 0 1

0

5

10

15

20

25

30

w

Dep

th (

m)

a

0 0.5 1<−u’w’>U

z

b

−0.5 0 0.5−g ρ

0−1 <ρ’w’>

c

−0.5 0 0.5−ρ

0−1<p’w’>

z

d

Figure 4.13. Terms in the kinetic energy equation for the unstable mode

responsible for the high frequency oscillations during period 1900. (a) Complex

vertical velocity (magnitude normalized to 1 in solid line and phase divided by π

in dashed line). (b) Shear production of perturbation kinetic energy, (c) work done

by gravity, and (d) vertical flux. Shear production, work and vertical flux are

divided by the maximum shear production. Horizontal solid line indicates the

critical level.

Page 171: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Shear generated high-frequency waves 147

0 0.02 0.04

0

5

10

15

20

25

30

w’ (m s−1)

Dep

th (

m)

a

0 0.1 0.2ρ’ kg (m−3)

b

997 998 999 1000ρ (kg m−3)

c

Figure 4.14. (a) Vertical velocity due to the unstable mode IV in Fig. 4.12e after

it has grown from an initial (arbitrary) maximum amplitude of 0.02 m s-1 at 4 m

depth during 1000 s until it reached a maximum amplitude of 0.05 m s-1. (b)

Associated density perturbations. (c) Total density obtained by adding the

background density in Fig. 4.8a and the density perturbation in Fig. 4.14b.

Page 172: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

148 Chapter 4

Page 173: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

149

Chapter 5

Conclusions

5.1. Summary

This thesis contributes to the understanding of the energy cascade in

stratified lakes by identifying the effects of bathymetry on the structure of the

basin’s natural internal modes, illustrating how several types of high-frequency

internal waves contribute in different ways to the heterogeneous distribution of

buoyancy flux, and by identifying an undocumented element in the energy flux

path.

In regards to basin-scale internal waves, this is the first study that

considers the combined effects of continuous stratification and real basin shape to

investigate the three-dimensional structure of the natural modes of oscillation in a

basin where the Earth’s rotation influences the motion. It was found that, in

addition to modulating the amplitude of azimuthal mode one natural modes

around the perimeter, the horizontal shape of the basin at the level of the

thermocline controls the generation of secondary cells in natural modes; natural

modes have several cells if each cell satisfies the dispersion relationship for

circular or elliptical basins. A quick prediction of the horizontal structure of

natural modes with multiple cells in irregular basins can now be made from the

dispersion relationship for circular or elliptical basins by fitting circles or ellipses

of the right dimensions into the contour of the lake at the level of the thermocline.

The sloping bottom produces localized intensification of the amplitude of

the natural modes near the bed due to focusing of internal wave rays over near-

critical slopes. This result supports and extends what was previously proposed for

Page 174: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

150 Chapter 5

two-dimensional basins. The locations where the amplitude of the basin-scale

internal waves is large are hot spots for the generation of processes, such as

increased dissipation or transference of energy to high frequency waves, that take

energy from the basin-scale waves to smaller scales. This intensification of the

motion also affects directly the water quality due to a higher resuspension of

sediments due to the increased shear over the bed.

Chapter 3 provides field evidence that summarizes the different types of

effects that high-frequency internal waves may have on mixing. Shear-generated

high-frequency internal waves dissipate quickly, having only a local and

immediate effect in vertical mixing. In contrast, solitary waves generated by

steepening of basin-scale waves may travel long distances without generating

significant mixing; they may experience fission at the interaction with a sill but

only contribute significantly to mixing when they break at the end of the lake.

Hence, they produce mixing far from the location where they are generated.

Although the behaviour of an internal bore depends on its strength, in general,

they are able to produce strong vertical mixing for long distances as they

propagate. For a lake with frequent formation of internal bores, the generalized

idea that mixing in the benthic boundary layer is much larger than at the interior

of the lake needs to be reviewed.

A major result of this research is the discovery of an unidentified

mechanism that participates in two ways in the energy flux path. Shear introduced

by the wind in the surface layer generates a shear-unstable mode with a vertical

structure that produces vertical transport of the kinetic energy in the mode. Part of

the energy returns back to the mean flow in the metalimnion due to the interaction

with the shear generated by the basin-scale waves. The density fluctuations

Page 175: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Conclusions 151

associated with the unstable mode strain the density field in the metalimnion and

trigger secondary instabilities. These secondary instabilities generate special

turbulent events in the metalimnion (Yeates et al. 2007). Hence, the shear-

unstable mode extracts energy from the surface shear, produces localized mixing

in the surface layer and in the metalimnion, and energizes a metalimnetic jet.

5.2. Future work

The technique I developed to investigate the spatial structure of basin-

scale internal waves has limitations. For instance, it is difficult to identify the

signature of high vertical mode waves due to the initial condition utilized. In

addition, the separation of modes when different modes have similar periods is

not straightforward. An extension of the existing eigenvalue numerical models for

two-dimensional basins with sloping bottom and continuous stratification to three

dimensions becomes computationally expensive for the numerical techniques and

computational power available. I think, however, that an eigenvalue numerical

model to calculate the three-dimensional structure of the natural modes needs to

be developed. The frequent improvements in computational power and numerical

techniques will facilitate this. Such a model would be very useful to investigate

which natural modes are favourably excited by winds with known period and

spatial structure.

The detailed calculation of the generation, propagation and dissipation of

high frequency waves requires the utilization of non-hydrostatic models with high

spatial and temporal resolution, which is computationally very expensive even for

a small lake. An alternative to overcome this is to parameterize those processes

for the different types of high-frequency waves into coarse grid models, which is

Page 176: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

152 Chapter 5

harder when the sink of wave energy is located away from the source. Another

alternative is to use adaptable grid models able to refine the grid wherever is

necessary to track a small-scale process.

Page 177: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

153

Bibliography

Afanasyev, Ya. D. and W. R. Peltier. 2001. On breaking internal waves over the

sill in Knight inlet. Proc. R. Soc. Lond A. 457: 2799-2825.

Amorocho, J., and J. J. DeVries. 1980. A new evaluation of the wind stress

coefficient over water surfaces. J. Geophys. Res. 85: 433-442.

Antenucci, J. P., J. Imberger, and A. Saggio. 2000. Seasonal evolution of the

basin-scale internal wave field in a stratified lake. Limnol. Oceanogr. 45: 1621-

1638.

Antenucci, J. P., and J. Imberger. 2001a. Energetics of long internal gravity waves

in large lakes. Limnol. Oceanogr. 46: 1760-1773.

Antenucci, J. P. and J. Imberger. 2001b. On internal waves near the high-

frequency limit in an enclosed basin. J. Geophys. Res. 106: 22465-22474.

Antenucci, J. P. and J. Imberger. 2003. The seasonal evolution of wind/internal

wave resonance in Lake Kinneret. Limnol. Oceanogr. 48: 2055-2061.

Apel, J. R., Holbrock, J. R., Liu, A. K., and Tsai, J. J. 1985. The Sulu Sea internal

soliton experiment. J. Phys. Oceanogr. 15: 1625-1651.

Page 178: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

154 Bibliography

Appt, J., J. Imberger, and H. Kobus. 2004. Basin-scale motion in stratified Upper

Lake Constance. Limnol. Oceanogr. 49: 919-933.

Appt, J. and S. Stumpp. 2002. Die Bodensee-Messkampagne 2001. IWS/CWR

Lake Constance Measurement Program 2001. Mitt.heft 111. Institut für

Wasserbau. Universität Stuttgart.

Bäuerle, E. 1998. Excitation on internal seiches by periodic forcing, p. 167-178.

In J. Imberger [ed.], Physical processes in lakes and oceans. Coastal and Estuarine

Studies. V. 54. AGU.

Bäuerle, E., D. Ollinger, and J. Ilmberger. 1998. Some meteorological,

hydrological, and hydrodynamical aspects of Upper Lake Constance. Arch.

Hydrobiol. Spec. Issues Adv. Limnol. 53: 31-83.

Beletsky, D., W. P. O’Connor, D. J. Schwab, and D. E. Dietrich. 1997. Numerical

simulation of internal Kelvin waves and coastal upwelling fronts. J. Phys.

Oceanogr. 27: 1197-1215.

Bendat, J., and A. Piersol. 1986. Random data. Analysis and measurement

procedures. 2nd ed. Wiley.

Bennett, J. R. 1973. A theory of large amplitude Kelvin waves. J. Phys. Oceanogr.

3: 57-60.

Page 179: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Bibliography 155

Bennett, J. R. 1977. A three-dimensional model of Lake Ontario’s summer

circulation. I: Comparison with observations. J. Phys. Oceanogr. 7: 591-601.

Benney, D. J. 1966. Long nonlinear waves in fluid flows. J. Math. Phys. 45: 52-

69.

Boegman, L., J. Imberger, G. N. Ivey, and J. P. Antenucci. 2003. High frequency

internal waves in large stratified lakes. Limnol. Oceanogr. 48: 895-919.

Boegman, L., G. N. Ivey, and J. Imberger. 2005a. The energetics of large-scale

internal wave degeneration in lakes. J. Fluid Mech. 531: 159-180.

Boegman, L., G. N. Ivey, and J. Imberger. 2005b. The degeneration of internal

waves in lakes with sloping topography. Limnol. Oceanogr. 50: 1620-1637.

Boehrer, B. 2000. Modal response of a deep stratified lake: western lake

Constance. J. Geophys. Res. 105: 28837-28845.

Caulfield, C. P. 1994. Multiple linear instability of layered stratified shear flow. J.

Fluid Mechanics. 258: 255-285.

Csanady, G. T. 1967. Large-scale motion in the Great Lakes. J. Geophys. Res. 72:

4151-4162.

Page 180: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

156 Bibliography

Csanady, G. T. 1972. Response of large stratified lakes to wind. J. Phys.

Oceanogr. 2: 3-13.

Csanady, G. T. 1976. Topographic waves in Lake Ontario. J. Phys. Oceanogr. 6:

93-103.

Csanady, G. T. 1977. Intermittent “full” upwelling in Lake Ontario. J. Geophys.

Res. 82: 397-419.

Csanady. G. T. 1982. On the structure of transient upwelling events. J. Phys.

Oceanogr. 12: 84-96.

Davis, P. A. and W. R. Peltier. 1976. Resonant parallel shear instability in the

stably stratified planetary boundary layer. J. Atmos. Sci. 33: 1287-1300.

Dauxois, T., and W.R. Young. 1999. Near-critical reflection of internal waves. J.

Fluid Mech. 390: 271-295.

De Bass, A. F. and A. G. M. Driedonks. 1985. Internal gravity waves in a stably

stratified boundary layer. Boundary Layer Meteorol. 31: 303-323.

De Silva, I. P. D., J. Imberger, and G. N. Ivey. 1997. Localised mixing due to a

breaking internal wave ray at a sloping bed. J. Fluid Mech. 350: 1-27.

Page 181: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Bibliography 157

Dintrans, B., M. Rieutord, and L. Valdettaro. 1999. Gravito-inertial waves in a

rotating stratified sphere or spherical shell. J. Fluid Mech. 398: 271-297.

Drazin, P. G. 1958. The stability of a shear layer in an unbounded heterogeneous

inviscid fluid. J. Fluid Mech. 4: 214-224.

Drazin, P. G. and R. S. Johnson. 1989. Solitons: an introduction. Cambridge

University Press.

Drazin, P. G. and W. H. Reid. 1981. Hydrodynamic stability. Cambridge

University Press.

Eriksen, C.C. 1982. Observations of internal wave reflections off sloping bottoms.

J. Geophys. Res. 87: 525-538.

Farmer, D. M. 1978. Observations of long nonlinear internal waves in a lake. J.

Phys. Oceanogr. 8: 63-73.

Farmer, D. M. and J. D. Smith. 1980. Tidal interaction of stratified flow with a sill

in Knight inlet. Deep-sea Res. 27A: 239-254.

Fricker, P. D., and H. M. Nepf. 2000. Bathymetry, stratification and internal

seiche structure. J. Geophys. Res. 105: 14237-14251.

Page 182: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

158 Bibliography

Gan, J. and G. Ingram. 1992. Internal hydraulics, solitons and associated mixing

in a stratified sound. J. Geophys. Res. 97: 9669-9688.

Gloor, M., A. Wüest, and M. Münnich. 1994. Benthic boundary mixing and

resuspension induced by internal seiches. Hydrobiologia. 284: 59-68.

Gómez-Giraldo, A., J. Imberger, and J. P. Antenucci. 2006. Spatial structure of

the dominant basin-scale internal waves in Lake Kinneret. Limnol. Oceanogr. 51:

229-246. (Chapter 2 of this thesis).

Gómez-Giraldo, A., J. Imberger, J. P. Antenucci, and P. S. Yeates. 2007.Wind-

shear generated high-frequency internal waves as precursors to mixing in a

stratified lake. Accepted for publication in Limnol. Oceanogr. (Chapter 4 of this

thesis).

Gonella, J. 1972. A rotary-component method for analysing meteorological and

oceanographic vector time series. Deep-Sea Res. 19: 833-846.

Gómez-Giraldo, A., J. Imberger, and J. P. Antenucci. 2006. Spatial structure of

the dominant basin-scale internal waves in Lake Kinneret. Limnol. Oceanogr. 51:

229-246.

Goudsmit, G. H., F. Peeters, M. Gloor, and A. Wüest. 1997. Boundary versus

internal diapycnal mixing in stratified natural waters. J. Geophys. Res. 102:

27,903-27,914.

Page 183: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Bibliography 159

Hamblin, P. F. 1977. Short-period internal waves in the vicinity of a river-induced

shear zone in a fiord lake. J. Geophys. Res. 82: 3167-3174.

Hazel, P. 1972. Numerical studies of the stability of inviscid stratified shear flows.

J. Fluid Mech. 51: 39-61.

Heaps, N. S., and A. E. Ramsbottom. 1966. Wind effects on water in a narrow

two-layered lake. Phil. Trans. R. Soc. London, ser A. 259: 391-430.

Helfrich, K. R., 1992. Internal solitary wave breaking and run-up on a uniform

slope. J. Fluid Mech. 243: 133-154.

Helfrich, K. R. and W. K. Melville. 1986. On long nonlinear waves over slope-

shelf topography. J. Fluid Mech. 167: 285-308.

Hicks, B. B. 1975. A procedure for the formulation of bulk transfer coefficients

over water. Bound. Layer Meteorol. 8: 515-524.

Hodges, B. R., J. Imberger, A. Saggio, and K. B. Winters. 2000. Modeling basin-

scale internal waves in a stratified lake. Limnol. Oceanogr. 45: 1603-1620.

Hogg, A. M., K. B. Winters, and G. N. Ivey. 2001. Linear internal waves and the

control of stratified exchange flows. J. Fluid Mech. 447: 357-375.

Page 184: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

160 Bibliography

Holloway, P. E., E. Pelinovsky, T. Talipova, and B. Barnes. 1997. A nonlinear

model of internal tide transformation on the Australian North West Shelf. J. Phys.

Oceanogr. 27: 871-896.

Horn. D. A., L. G. Redekopp, J. Imberger, and G. N. Ivey. 2000. Internal wave

evolution in a space-time varying field. J. Fluid Mech. 424: 279-301.

Horn, D. A., J. Imberger, and G. N. Ivey. 2001. The degeneration of lage-scale

interfacial gravity waves in lakes. J. Fluid. Mech. 434: 181-207.

Horn, W., C. H. Mortimer, and D. J. Schwab. 1986. Wind-induced internal

seiches in Lake Zurich observed and modelled. Limnol. Oceanogr. 31: 1232-

1254.

Howard, L. N. 1961. Note on a paper by John W. Miles. J. Fluid Mech. 10: 509-

512.

Howard, L. N. and S. A. Maslowe. 1973. Stability of stratified shear flows.

Boundary Layer Meteorol. 4: 511-523.

Hunkins, K. and M. Fliegel. 1973. Internal undular surges in Seneca Lake: A

natural occurrence of solitons. J. Geophys. Res. 78: 539-548.

Page 185: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Bibliography 161

Hutter, K., G. Bauer, Y. Wang, and P. Güting. 1998. Forced motion response in

enclosed lakes, p. 137-166. In J. Imberger [ed.], Physical processes in lakes and

oceans. Coastal and Estuarine Studies. V. 54. AGU.

Imberger, J. 1998. Flux paths in a stratified lake: A review, p. 1-17. In J. Imberger

[ed.], Physical processes in lakes and oceans. Coastal and Estuarine Studies. V.

54. AGU.

Imberger, J., and J.C. Patterson. 1981. A dynamic reservoir simulation model-

DYRESM 5, p. 310-361. In H. Fisher [ed.], Transport models for inland and

coastal waters. Academic Press.

Imberger, J. and R. Head. 1994. Measurement of turbulent properties in a natural

system. Fundamentals and advancements in hydraulic measurements and

experimentation. Am. Soc. of Civ. Eng. New York.

Imboden, D. M. 1990. Mixing and transport in lakes: Mechanisms and ecological

relevance. p. 47-80. In M. M. Tilzer and C. Serruya [ed.]. Large Lakes: Ecological

structure and function. Springer-Verlag.

Ivey. G. N., and T. Maxworthy. 1992. Mixing driven by internal Kelvin waves in

large lakes and the coastal oceans. Proceedings 11th Australasian Fluid Mechanics

Conference, Hobart, Australia. 303-306.

Page 186: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

162 Bibliography

Jacquet, J. 1983. Simulation of the thermal regime of rivers, p. 150-176. In G. T.

Orlob [ed.], Mathematical modelling of water quality: Streams, lakes and

reservoirs. Wiley-Interscience.

Kaiser, G. 1994. A friendly guide to wavelets. Birkhauser.

Kistovich, Y. V., and Y. D. Chashechkin. 1995. The reflection of beams of

internal gravity waves at a flat rigid surface. J. Appl. Maths. Mechs. 59: 579-585.

Klemp, J. B. and D. K. Lilly. 1975. Dynamics of wave-induced downslope winds.

J. Atmos. Sci. 32: 320-339.

Klemp J. B., R. Rotunno, and W. C. Skamarock. 1997. On the propagation of

internal bores. J. Fluid Mech. 331: 81-106.

Kundu, P. K., and I. M. Cohen. 2002. Fluid Mechanics. 2nd ed. Academic Press.

Lalas, D. P. and F. Einaudi. 1976. On the characteristics of gravity waves

generated by atmospheric shear layers. J. Atmos. Sci. 33: 1248-1259.

Lamb, K. G. 1997. Particle transport by nonbreaking, solitary internal waves, J.

Geophys. Res. 102: 18641-18660.

Lamb, K. G. 2002. A numerical investigation of solitary internal waves with

trapped cores formed via shoaling. J. Fluid Mech. 451: 109-144.

Page 187: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Bibliography 163

Lamb, K. G. 2004. Boundary-layer separation and internal wave generation at the

Knight inlet sill. Proc. R. Soc. Lond. A. 460: 2305-2337.

Laval, B., B. R. Hodges, and J. Imberger. 2003a. Numerical diffusion in 3D,

hydrostatic, z-level lake models. J. Hydraulic Eng. 129: 215-224.

Laval, B., J. Imberger, B. R. Hodges, and R. Stocker. 2003b. Modeling circulation

in lakes: Spatial and temporal variations. Limnol. Oceanogr. 48: 983-994.

Lawrence, G. A., F. K. Browand, and L. G. Redekopp. 1991. The stability of a

sheared density interface. Phys. Fluids A. 3: 2360-2370.

LeBlond, P. H., and L. A. Mysak. 1978. Waves in the ocean. Elsevier

Lemckert, C. and J. Imberger. 1998. Turbulent benthic boundary layer mixing

events in fresh water lakes, p. 503-516. In J. Imberger [ed.], Physical processes in

lakes and oceans. Coastal and Estuarine Studies. V. 54. AGU.

Lemmin, U. 1987. The structure and dynamics of internal waves in Baldeggersee.

Limnol. Oceanogr. 32: 43-61.

Lemmin, U., C. H. Mortimer, and E. Bäuerle. 2005. Internal seiche dynamics in

Lake Geneva. Limnol. Oceanogr. 50: 207-216.

Lighthill, J. 1978. Waves in Fluids. Cambridge.

Page 188: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

164 Bibliography

Luketina, D. and J. Imberger. 1989. Turbulence and Entrainment in a Buoyant

Surface Plume. J. Geophys. Res. 94: 12619-12636.

Maas, L. R. M. 2001. Wave focusing and ensuing mean flow due to symmetry

breaking in rotating fluids. J. Fluid Mech. 437: 13-28.

Maas, L. R. M., and F.-P. A. Lam. 1995. Geometric focusing of internal waves. J.

Fluid Mech. 300: 1-41.

Maas, L. R. M., D. Benielli, J. Sommeria, and F.-P. A. Lam. 1997. Observation of

an internal wave attractor in a confined, stably stratified fluid. Nature. 388: 557-

561.

MacIntyre, S., K. M. Flynn, R. Jellison, and J. R. Romero. 1999. Boundary

mixing and nutrient fluxes in Mono Lake, California. Limnol. Oceanogr. 44: 512-

529.

Manders, A. M. M., and L. R. M. Maas. 2003. Observations of inertial waves in a

rectangular basin with in sloping boundary. 493: 59-88.

Manders, A. M. M., and L. R. M. Maas. 2004. On the three-dimensional structure

of the inertial wave field in a rectangular basin with one sloping boundary. Fluid

Dyn. Res. 35: 1-21.

Page 189: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Bibliography 165

Marti, C. L. 2004. Exchange processes between littoral and pelagic waters in a

stratified lake, PhD thesis. University of Western Australia.

Michallet, H and Ivey, G. N. 1999. Experiments on mixing due to internal solitary

waves breaking on uniform slopes. J. Geophys. Res. 104: 13467-13477.

Miles, J. W. 1961. On the stability of heterogeneous shear flow. J. Fluid Mech.

10: 496-508.

Miles, J. W. and L. N. Howard. 1964. Note on a heterogeneous shear flow. J.

Fluid Mech. 20: 311-313.

Monismith, S. G. 1985. Wind-forced motions in stratified lakes and their effect on

mixed-layer shear. Limnol. Oceanogr. 30: 771-783.

Monismith, S. G. 1986. An experimental study of the upwelling response of

stratified reservoirs to surface shear stress. J. Fluid Mech. 171: 407–439.

Monismith, S. 1987. Modal response of Reservoirs to wind stress. J. Hydr. Eng.

113: 1290-1306.

Mortimer, C. H., D. C. McNaught, and K. M. Stewart. 1968. Short internal waves

near their high-frequency limit in central Lake Michigan. Proc. 11th Conf. Great

Lakes Res. II, 454-469.

Page 190: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

166 Bibliography

Mortimer, C. H. 1974. Lake Hydrodynamics. Verh. Int. Ver. Limnol. 20: 124-

197.

Mortimer, C. H. and W. Horn. 1982. Internal wave dynamics and their

implications for plankton biology in the Lake of Zurich. Vierteljahrsschr.

Naturforsch. Ges. Zurich. 127: 299-318.

Münnich, M., A. Wüest, and D. M. Imboden. 1992. Observations of the second

vertical mode of the internal seiche in an alpine lake. Limnol. Oceanogr. 37:

1705-1719.

Münnich, M. 1996. The influence of bottom bathymetry on internal seiches in

stratified media. Dyn. Atmos. Oceans. 23: 257-266.

Norton, M. P. 1989. Fundamentals of noise and vibration analysis for engineers.

Cambridge.

Osborne, A. R. and Burch, T. I. 1980. Internal solitons in the Andaman Sea.

Science. 208: 451-469.

Ostrovsky, I., Y. Z. Yacobi, P. Walline, and I. Kalikhman. 1996. Seiche-induced

mixing: its impact on lake productivity. Limnol. Oceanogr. 41: 323-332.

Page 191: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Bibliography 167

Ou, H., and J. R. Bennett. 1979. A theory of the mean flow driven by long internal

waves in a rotating basin, with application to Lake Kinneret. J. Phys. Oceanogr. 9:

1112-1125.

Phillips, O. M. 1966. The dynamics of the upper ocean. Cambridge University

Press.

Pineda, J. 1999. Circulation and larval distribution in internal tidal bore warm

fronts. Limnol. Oceanogr. 44: 1400-1414.

Rhines, P. 1970. Edge-, Bottom-, and Rossby waves in a rotating stratified fluid.

Geophys. Fluid Dyn. 1: 273-302.

Rieutord, M, B. Georgeot, and L. Valdettaro. 2001. Inertial waves in a rotating

spherical shell: attractors and asymptotic spectrum. J. Fluid Mech. 435: 103-144.

Rottman, J. W. and J. E. Simpson. 1989. The formation of internal bores in the

atmosphere: A laboratory model. Q. J. R. Meteorol. Soc. 115: 941-963.

Rueda, F. J., S. G. Schladow, and S. O. Palmarsson. 2003. Basin-scale internal

wave dynamics during a winter cooling period in a large lake. J. Geophys. Res.

108: art. no. 3097.

Saggio, A., and J. Imberger. 1998. Internal wave weather in a stratified lake.

Limnol. Oceanogr. 43: 1780-1795.

Page 192: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

168 Bibliography

Saggio, A. and J. Imberger .2001. Mixing and turbulent fluxes in the metalimnion

of a stratified lake. Limnol. Oceanogr. 46: 392-409.

Salvadè, G., F. Zamboni, and A. Barbieri. 1988. Three-layer model of the North

basin of the lake of Lugano. Annales Geophysicae. 6: 463-474.

Serruya, S. 1975. Wind, water temperature and motions in Lake Kinneret: general

pattern. Verh. Int. Ver. Limnol. 19: 73-87.

Schwab, D. J. 1977. Internal free oscillation in Lake Ontario. Limnol. Oceanogr.

22: 700-708.

Schwab, D. J., and D. Beletsky. 1998. Propagation of Kelvin waves along

irregular coastlines in finite-difference models. Advances in water resources. 22:

239-245.

Smyth, W. D. and W. R. Peltier. 1990. Three-dimensional primary instabilities of

a stratified, dissipative, parallel flow. Geophys. Astrophys. Fluid Dyn. 52: 249-

261.

Spigel, R. H. and J. Imberger. 1980. The classification of mixed-layer dynamics in

lakes of small to medium size. J. Phys. Oceanogr. 10: 1104-1121.

Stevens, C. L. 1999. Internal waves in a small reservoir. J. Geophys. Res. 104:

15777-15788.

Page 193: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Bibliography 169

Stocker, R., and J. Imberger. 2003a. Energy partitioning and horizontal dispersion

in a stratified rotating lake. J. Phys. Oceanogr. 33: 512-529.

Stocker, R., and J. Imberger. 2003b. Horizontal transport and dispersion in the

surface layer of a medium size lake. Limnol. Oceanogr. 48: 971-982.

Stocker, T., and K. Hutter. 1987. Topographic waves in channels and lakes on the

f-plane. Lecture notes on Coastal and Estuarine Studies. Springer.

Sun, C., W. D. Smyth, and J. N. Moum. 1998. Dynamic instability of stratified

shear flow in the upper equatorial Pacific. J. Geophys Res. 103: 10323-10337.

Sutherland, B. R. 1996. Dynamic excitation on internal gravity waves in the

Equatorial oceans. J. Phys. Oceanogr. 26: 2398-2419.

Sutherland, B. R. and W. R. Peltier. 1994. Turbulence transition and internal wave

generation in density stratified jets. Phys. Fluids. 6: 1267-1284.

Sutherland, B. R. and W.R. Peltier. 1995. Internal gravity wave emission into the

middle atmosphere from a model tropospheric jet. J. Atmos. Sci. 52: 3214-3235.

Taylor, G. I. 1931. Effect of variation in density on the stability of superposed

streams of fluid. Proc. R. Soc. Lond. A 132: 499-523.

Page 194: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

170 Bibliography

Taylor, J. R. 1993. Turbulence and mixing in the boundary layer generated by

shoaling internal waves. Dyn. Atmos. Oceans. 19: 233-258.

Thorpe, S. A. 1968. A method of producing a shear flow in a stratified fluid. J.

Fluid Mech. 32: 693-704.

Thorpe, S. A. 1971. Experiments on the instability of stratified shear flows:

miscible fluids. J. Fluid Mech. 46: 299-319.

Thorpe, S. A. 1973. Turbulence in stratified fluids: A review of laboratory

experiments. Boundary Layer Meteor. 5: 95-119.

Thorpe, S. A. 1974. Near-resonant forcing in a shallow two-layer fluid: a model

for the internal surge in Loch Ness? J. Fluid Mech. 63: 509-527.

Thorpe, S. A., A. Hall, and I. Crofts. 1972. The internal surge in Lock Ness.

Nature. 237: 96-98.

Thorpe, S. A., J. M. Keen, R. Jiang, and U. Lemmin. 1996. High-frequency

internal waves in Lake Geneva. Philos. Trans. R. Soc. London. Ser A. 354: 237-

257.

Torrence, C. and G. P. Compo. 1998. A practical guide to wavelet analysis. Bull.

Am. Meteorol. Soc. 79: 61-78.

Page 195: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

Bibliography 171

Turner, J. S. 1973. Buoyancy effects in fluids. Cambridge.

Vlasenko, V. I. and K. Hutter. 2002. Transformation and disintegration of

strongly nonlinear internal waves by topography in stratified lakes. Annales

Geophysicae. 20: 2087-2103.

Wang, Y., K. Hutter, and E. Bäuerle. 2000. Wind-induced baroclinic response of

Lake Constance. Annales Geophysicae. 18: 1488-1501.

Wake, G. W., G. N. Ivey, and J. Imberger. 2005. The temporal evolution of

baroclinic basin-scale waves in a rotating circular basin. J. Fluid Mech. 523: 367-

392.

Wetzel, R. G. 2001. Limnology: lake and river ecosystems. Academic Press.

Wiegand, R. C. and E. C. Carmack. 1986. The climatology of internal waves in a

deep temperate lake. J. Geophys. Res. 91: 3951-3958.

Woods, J. D. 1968. Wave-induced shear instability in the summer thermocline. J.

Fluid Mech. 32: 791-800.

Wüest, A., G. Piepke, and D.C. Senden. 2000. Turbulent kinetic energy balance as

a tool for estimating vertical diffusivity in wind-forced stratified waters. Limnol.

Oceanogr. 45: 1388-1400.

Page 196: Observations of energy flux mechanisms associated with internal waves · Observations of energy transfer mechanisms associated with internal waves Evelio Andrés Gómez Giraldo B.Eng

172 Bibliography

Yeates, P. S., and J. Imberger. 2003. Pseudo two-dimensional simulations of

internal and boundary fluxes in stratified lakes and reservoirs. Intl. J. River Basin

Management 1: 297-319.

Yeates, P. S., J. Imberger, and A. Gómez-Giraldo. 2007. Turbulent mixing

intensity and its relationship to the gradient Richardson number in the water

column of a stratified lake. Accepted for publication in Limnol. Oceanogr.