Thesis Kenji Shimizu

download Thesis Kenji Shimizu

of 143

Transcript of Thesis Kenji Shimizu

  • 8/2/2019 Thesis Kenji Shimizu

    1/143

    Application of modal analysis

    to strongly stratified lakes

    Kenji Shimizu

    B. Eng. (Mechanical) Tokyo Institute of Technology, Tokyo, Japan.

    M. Eng. (Environmental) Tokyo Institute of Technology, Tokyo, Japan.

    This thesis is presented for the degree of Doctor of Philosophyof The University of Western Australia.

    April 2008

  • 8/2/2019 Thesis Kenji Shimizu

    2/143

  • 8/2/2019 Thesis Kenji Shimizu

    3/143

    iii

    Abstract

    Modal analysis for strongly stratified lakes was extended to obtain a better

    understanding of the dynamics of the basin-scale motions. By viewing the basin-scale

    motions as a superposition of modes, that have distinct periods and three-dimensionalstructures, the method provides a conceptual understanding for the excitation, evolution,

    and damping of the basin-scale motions. Once the motion has been decomposed into

    modes, their evolution and energetics may be extracted from hydrodynamic simulation

    results and field data. The method was applied to Lake Biwa, Japan, and Lake Kinneret,

    Israel, and used for a theoretical study.

    The real lake applications showed that winds excited basin-scale motions that had a

    surface layer velocity structure similar to the wind stress pattern. Three-dimensionalhydrodynamics simulations of Lake Biwa indicated that most of the energy input from

    winds was partitioned into the internal waves that decayed within a few days. The gyres,

    on the other hand, received much less energy but dominated the dynamics during calm

    periods due to their slow damping. Analyses of field data from Lake Kinneret suggested

    that the internal waves, excited by the strong winds every afternoon, were damped over

    a few days primarily due to bottom friction. Theoretical investigations of damping

    mechanisms of internal waves revealed that bottom friction induced a velocity anomalyat the top of the boundary layer that drained energy from the nearly inviscid interior by

    a combination of internal wave cancelling and spin-down.

    These results indicate that gyres induce long-term horizontal transport near the

    surface and internal waves transfer energy from winds to near-bottom mixing. Modal

    structure of dominant basin-scale internal waves can induce large heterogeneity of near-

    bottom mass transfer processes. The method presented here provides a tool to determine

    how basin-scale motions impact on biogeochemical processes in stratified lakes.

  • 8/2/2019 Thesis Kenji Shimizu

    4/143

    iv

  • 8/2/2019 Thesis Kenji Shimizu

    5/143

    v

    Table of contents

    Abstract ............................................................................................................................iii

    Table of contents ............................................................................................................... v

    List of tables ....................................................................................................................vii

    List of figures .................................................................................................................viii

    List of symbols ...............................................................................................................xiii

    Acknowledgements.......................................................................................................xvii

    Preface............................................................................................................................xix

    Chapter 1. Introduction ..................................................................................................... 1

    1.1 Motivation..........................................................................................................1

    1.2 Overview............................................................................................................2

    Chapter 2. Horizontal structure and excitation of primary motions in a stronglystratified lakes..........................................................................................................4

    2.1 Introduction........................................................................................................4

    2.2 Field Data...........................................................................................................8

    2.3 Theory of modal analysis .................................................................................10

    2.3.1 Basic equations and solutions .........................................................10

    2.3.2 Some property of modes .................................................................13

    2.3.3 Decoupled dynamic equations of individual modes ....................... 14

    2.4 Modal analysis of Lake Biwa........................................................................... 162.4.1 Numerical method........................................................................... 16

    2.4.2 Horizontal structure and excitation of internal waves.....................18

    2.4.3 Spatial structure and excitation of gyres .........................................20

    2.5 Three-dimensional hydrodynamic simulations of Lake Biwa ......................... 23

    2.5.1 Hydrodynamic model...................................................................... 23

    2.5.2 Preferential excitation of internal waves.........................................23

    2.5.3 Excitation of gyres by wind stress curl ...........................................26

    2.6 Discussion........................................................................................................27Appendix 2.A -- Self-adjointness of operator K ....................................................... 33

    Chapter 3. Energetics and damping of internal waves in a strongly stratified lake....... 37

    3.1 Introduction......................................................................................................37

    3.2 Theory of modal analysis with linear damping................................................ 41

    3.2.1 Shallow water equations and the associated modes ........................ 41

    3.2.2 Governing equations of modal amplitude and energy ....................45

    3.3 Study site and field data ...................................................................................46

    3.3.1 Study site .........................................................................................46

  • 8/2/2019 Thesis Kenji Shimizu

    6/143

    vi

    3.3.2 Field data.........................................................................................46

    3.4 Spatial structure of basin-scale internal waves in Lake Kinneret ....................49

    3.5 Energetics and damping of internal waves ...................................................... 53

    3.5.1 Extraction of internal waves by mode fitting.................................. 53

    3.5.2 Energetics of basin-scale internal waves ........................................ 583.5.3 Damping of basin-scale internal waves by bottom friction ............ 60

    3.6 Spatial variability of near-bottom transport processes..................................... 61

    3.7 Discussion ........................................................................................................ 64

    Appendix 3.A -- Derivation of modal equations for a weakly dissipative system....68

    Chapter 4. Damping mechanisms of internal waves in a continuously stratifiedrotating basin ......................................................................................................... 71

    4.1 Introduction ...................................................................................................... 71

    4.2 Governing equations and scaling ..................................................................... 744.3 Ekman normal velocities induced by oscillatory boundary layers .................. 77

    4.3.1 Bottom boundary layer.................................................................... 77

    4.3.2 Sidewall boundary layer.................................................................. 80

    4.3.3 Bottom corner region ...................................................................... 82

    4.4 Perturbation analysis of damped internal waves.............................................. 83

    4.4.1 Inviscid solutions ............................................................................ 84

    4.4.2 Correction to the inviscid basin-scale internal waves due to

    boundary layer presence ......................................................................... 874.5 Internal wave cancelling in a flat-bottomed rectangular basin ........................ 90

    4.6 Damping of gravity waves in circular basins................................................... 93

    4.7 Discussion ...................................................................................................... 101

    4.8 Conclusions .................................................................................................... 103

    Appendix 4.A -- Orthogonality of modes in a continuously stratified rotatingbasin with arbitrary shape .............................................................................. 104

    Appendix 4.B -- Horizontal modes in a flat-bottomed circular basin .................... 108

    Appendix 4.C -- Horizontal modes in a flat-bottomed circular basin .................... 110Chapter 5. Conclusions ................................................................................................. 113

    5.1 Summary ........................................................................................................ 113

    5.2 Recommendations for future work ................................................................ 114

    Bibliography.................................................................................................................. 117

  • 8/2/2019 Thesis Kenji Shimizu

    7/143

    vii

    List of tables

    Table 2.1. Some characteristics of the internal waves and gyres in Lake Biwa. Some of thetopographic waves and surface waves (seiches) are also included to illustrate thespectrum of the modes. The surface and internal waves are numbered in decreasingorder of their natural periods, while the geostrophic gyres are numbered in the opposite

    order. No number is assigned to topographic waves. Abbreviations are HM:horizontal mode, PE: percentage of the available potential energy to the total energy,

    [ ]rE : mean modal total energy in the hydrodynamic simulation during BITEX 93, GS:

    geostrophic gyre (calculated as Rossby wave in the upper layer), TW: topographicwave, IW: (vertical mode 1) internal wave, SW: surface wave (seiche), CG: cyclonicgyre, AG: anticyclonic gyre, CC: cyclonic cell, AC: anticyclonic cell, NL: nodal line,NB: North Basin, N: north, S: south, and M: middle. Inertial period is 20.7 h. Calculated as the ratio of sum of the first two terms in Eq. 2.15 to sum of all the terms.# The sign [r] in the superscript indicates the sum of the conjugate pair, which havepositive and negative angular velocity with the same magnitude and the same modalstructure. References in the last column indicate the report of similar structure.Data sources are 1) Endoh (1986), 2) Endoh and Okumura (1993), 3) Kumagai et al.

    (1998), 4) Kanari (1975), 5) Saggio and Imberger (1998), 6) Endoh et al. (1995a), 7)Okamoto and Endoh (1995), and 8) Kanari (1974). All data except 8) are based on fieldobservation during summer, while 8) is based on a two layer model where the depths ofthe North and South Basin are taken as 50 m and 5 m, respectively........................................17

    Table 3.1. Characteristics of basin-scale internal waves in Lake Kinneret. Modal indices (firstcolumn) are assigned in ascending order of the angular frequency over all internalwaves (irrespective of their vertical structure), and the names (second column) areassigned on the basis of the corresponding internal wave structure in a flat-bottomedelliptical basin (see text). ( )r

    opt and ( )rb are damping rates estimated respectively by the

    mode-fitting technique and by using the equivalent linear bottom friction coefficient(Eq. 3.30). The estimated modal total energy ( [ ]rE ), rates of energy input ( [ ]rW ), and

    rates of energy dissipation ( [ ]rD ) averaged over the field campaign are shown in the lastthree columns. Abbreviations are V: vertical mode, H: horizontal mode, C: cyclonicwave, and A: anticyclonic wave. Numbers in parenthesis indicate angularfrequencies with bottom friction. Inertial frequency is 7.83 10-5 rad s-1. The middlevalues indicate the optimum damping rates and dissipation rates, and left and rightvalues show the sensitivity calculated based on 5% increase of the mean square error(Eq. 3.25). If the upper bound of ( )r

    opt goes to infinity, the corresponding upper bound of

    [ ]rD is calculated from the energy input assuming the energy is immediately dissipated. .......56

  • 8/2/2019 Thesis Kenji Shimizu

    8/143

    viii

    List of figures

    Fig. 2.1. Bathymetry of Lake Biwa and computational grids: (a) bathymetry of Lake Biwa andlocations of measurement stations, (b) magnified plan view of computational grid in A(common for both the modal analysis and ELCOM simulations), (c) computational gridalong cross-section B-B used for the modal analysis, and (d) schematic computational

    grid along cross-section B-B used for the ELCOM simulations (vertical grid resolutionis finer than the schematic). Thin and thick lines in panel (a) show bathymetry contourwith 10-m interval and horizontal boundaries of computational grid at the surface and18-m deep, respectively. The dashed line indicates a boundary between the northernand southern halves that are used for the calculation of spatially averaged wind stressand its curl (see Fig. 2.7)............................................................................................................9

    Fig. 2.2. Temperature profiles used for the modal analysis and the hydrodynamic simulations.The profiles in 1993 and 1994 were measured at BN50 and Sta. 17B, respectively. Forthe modal analysis with two-layer stratification, the depth of the interface was set to 18m (corresponds to maximum buoyancy frequency).................................................................10

    Fig. 2.3. Internal waves in Lake Biwa: (a) V1H1, (b) V1H2, (c) V1H4, and (d) V1H6. The

    middle panels correspond to the phase where the rates of energy input ( ( ))(RerW ) from

    spatially uniform winds are the largest ( )(r = 30, 120, 115, 70, see Fig. 2.4), and the

    right panels correspond to a quarter period after the left. Shading and vectors show theinterface displacement and the average velocity in the upper layer, respectively. In thesmaller left panels, shading and lines respectively show co-range contour and co-phaselines of wave trough with 45 interval. The co-phase lines with triangles correspond tothe phase shown in the middle figures, and the triangles show direction of rotation. Themodes are normalized to )(~ re = 2 GJ. Ellipses with dashed lines indicate cyclonic and

    anticyclonic cells used to calculate the dispersion relationship (see Fig. 2.11). ......................19

    Fig. 2.4. Normalized maximum energy input from spatially uniform winds blowing from

    different directions. Since energy input depends on the phase of the mode, themaximize energy input is normalized by the norm of )(~ r and f (thus vertical axis

    corresponds to ( ) ( ) ( ) 1/ , ,r r rW M f M f ). Wind direction is 0 when the wind is blowing

    from the north, and it is positive clockwise. ............................................................................ 20

    Fig. 2.5. Geostrophic gyres in Lake Biwa: (a) L1H1, (b) L1H2, and (c) L1H3. The panelsshow the most frequently observed phase during days 248-256 in the simulation results( )(r = -112.5, -135, 135). Shading and vectors show the interface displacement and

    the average velocity in the upper layer, respectively. The modes are normalized to )(~ re

    = 2 GJ......................................................................................................................................21

    Fig. 2.6. Typical wind patterns over Lake Biwa: (a) the SE wind, (b) the lake breeze, and (c)the NW wind. All examples were taken from the wind field during BITEX93. Winddata measured at five land-based stations were corrected against the offshore stationdata (see text) and linearly interpolated over the lake. Circles and a triangle indicateland-based and offshore meteorological stations, respectively................................................22

    Fig. 2.7. Spatially averaged (a) wind stress and (b) wind stress curl in 1994. Lake Biwa wassplit into northern and southern halves by the dashed line shown in Fig. 2.1, whichcorresponds to an approximate boundary between the First and Second Gyre. Datawere low-pass filtered with a cut-off frequency of 7 d. ........................................................... 22

    Fig. 2.8. Comparison of temperature profiles at BN50: (a) EW and NS components of thewind velocity (positive when wind is blowing from west and south, respectively), (b)the thermistor chain data, and (c) the results of the simulation. All data were averagedand subsampled at 30-min interval for plotting purposes. Please note that the

  • 8/2/2019 Thesis Kenji Shimizu

    9/143

    ix

    discrepancy of temperature near the surface is due to negligence of surface heatdynamics in the simulation.......................................................................................................24

    Fig. 2.9. Evolution of the amplitude and phase of (a) V1H1, (b) V1H2, (c) V1H4, and (d)V1H6 internal waves. Only relative magnitude of the amplitude have relevancebecause their absolute values depend on choice of normalization factor )(~ re . The phase

    increases at a rate of )(r

    when the modes are free, since the phase shown here is

    related to the modes with positive natural frequencies. The results during the shadedperiod are not reliable since the interface displacement was too large to apply the lineartheory. The phase in panels (c) and (d) are also not reliable when the amplitudes arelow, e.g., days 242-245.............................................................................................................25

    Fig. 2.10. Comparison of average velocity in the upper 10 m between field measurements byan ADCP over two days and the results of the simulations averaged over thecorresponding two days on (a) 22-23 June 1994 and (b) 26-27 July 1994. The observedvelocities along 11 transects in EW direction were linear interpolated to plot the ADCPdata. ..........................................................................................................................................26

    Fig. 2.11. Dispersion relationship of the fundamental mode (a) Kelvin wave and (b) Poincar

    wave in elliptic basins. Solid lines represent contour of the ratio of the natural angularfrequency to the inertial frequency 1 ( )rf , dotted lines indicate the natural period of

    the V1H1-H8 internal waves in Lake Biwa, symbols indicate dispersion relation ofeach cell in their horizontal structure (see Fig. 2.3), a and b are the half-length of themajor and minor axis of a basin or a cell, respectively, and

    longSand

    latS are the Burger

    number based on major and minor axis length respectively. For V1H1, the half-lengthof the thalweg was used as the major axis length.....................................................................28

    Fig. 2.12. Comparison of (a) the modal total energy [ ]rE (Eq. 2.20) and accumulated energy

    input ( )[ ]Re rW (Eq. 2.23), and (b) the phase deviation and accumulated phase input

    ( ) )()( 2/Im rr EW (Eq. 2.24) for V1H1 internal wave. In panel (b), the deviation of the

    phase was calculated by straightening the extracted phase (Fig. 2.9a) and subtractingthe constant increase of phase at a rate of )(r . Since the phase input was not reliable

    when the total energy was small, the accumulation of phase input was equated to thephase deviation on day 236.9 and 240.8 (triangles). See Table 2.1 and Fig. 2.9 for themeaning of [r] in the superscript and shaded period, respectively. ..........................................30

    Fig. 2.13. Comparison of the modal total energy [ ]rE (Eq. 2.20) and accumulated energy input

    ( )[ ]Re rW (Eq. 2.22) for (a) sum of all geostrophic modes, and (b) L1H1 and (c) L1H2

    geostrophic gyres. See Table 2.1 and Fig. 2.9 for the meaning of [r] in the superscriptand shaded period, respectively................................................................................................31

    Fig. 2.14. Energy partitioning among different classes of modes extracted from the results of

    the hydrodynamic simulation during BITEX93 using two layer approximation: (a)total energy [ ]rE and (b) accumulated energy input from winds ( )[ ]Re rW . The solid lines

    show the accumulation for each class, and dashed, dotted, and dot-dash linescorrespond to the accumulation up to the 1st, 5th, and 10th internal wave modes and 1st,20th, and 50th geostrophic modes. See Table 2.1 for the meaning of [r] in thesuperscript. ...............................................................................................................................32

    Fig. 3.1. Bathymetry of Lake Kinneret and locations of measurement stations. Circles andtriangles show locations of thermistor chains (equipped with wind sensor) and windsensors, respectively. Thick lines indicate horizontal boundaries of computational gridsfor modal computation with the interfaces at 14.4 m and 19.3 m deep, corresponding to25 C and 19 C isotherms. ......................................................................................................47

  • 8/2/2019 Thesis Kenji Shimizu

    10/143

    x

    Fig. 3.2. Results of the field experiment from 18 Jun to 01 Jul 2001. (a) Wind speed and (b)wind direction measured 2.4 m above the water surface at Ty (close to Tv), andisotherm depths at (c) Tg, (d) Tf, (e) T9, (f) T7, (g) Tv, and (h) T4. In panel (b), winddirection is 0 for wind blowing from north and positive clockwise. In panels (c-h),isotherms are plotted with 2 C interval with the top line being 25 C isotherm. 25 and19 C isotherms were used for the mode fitting, and dashed lines show basin-widetrends of the isotherm deepening used to calculate isotherm displacements. .......................... 48

    Fig. 3.3. Background vertical structures. (a) Profiles of temperature, , and the associatedbuoyancy frequency,N, at T4 averaged from day 170.5 to 171.5, (b) modaldisplacements, )( nh , and (c) modal horizontal velocity, )( n , induced by vertical mode 1

    ~ 3 internal waves under continuous stratification shown in panel (a), and (d) thicknessof the BBL, hBBL, taken from Fig. 5 in Marti and Imberger (2006). In order to calculatethe vertical modes, the bottom boundary condition was imposed at 27 m deep,corresponding to the average depth. Horizontal dotted lines indicate depths of theinterfaces used for modal analysis. .......................................................................................... 50

    Fig. 3.4. Diurnal internal waves in Lake Kinneret. (a, b) V2H3C and (c, d) V1H1C. Panels (a)and (c) correspond to internal wave field when net transport in the surface layer is from

    west to east, corresponding to the phase during the strong diurnal westerly winds, andpanels (b) and (d) a quarter period after panels (a) and (c). Shading in panels (a, c) and(b, d), respectively, indicates )325.0 + and )325.0 , which approximately

    correspond to V1 and V2 components of the displacements (see text). Solid linescorrespond to zero displacement, and vectors left and right, respectively, show layer-averaged velocity in the bottom and middle layer. .................................................................. 51

    Fig. 3.5. Same as Fig. 4 but for semi-diurnal internal waves. (a, b) V1H3C and (c, d) V1H1A........52

    Fig. 3.6. Comparison of 25 C and 19 C isotherm displacements at the six stations. (a) Tg,(b) Tf, (c) Tg, (d) T7, (e) Tv, and (f) T4. Solid lines: observation (detrended), dashedlines: prediction based on modal amplitude equations, Eq. 3.18 (used for fitting), anddotted lines: prediction (not used for fitting). .......................................................................... 55

    Fig. 3.7. Decomposition of predicted 25 C and 19 C isotherm displacements at Tf into fiveinternal wave components. (a) Sum of the five internal waves, (b) V2H3C, (c) V1H1C,(d) V2H1A, (e) V1H3C, and (f) V1H1A.................................................................................57

    Fig. 3.8. Estimated partitioning of (a) total (available potential plus kinetic) energy, (b)accumulated rate of work done by winds, and (c) accumulated rate of energydissipation. The lines show accumulation for each internal wave, so that the top linesshow the sum for the five internal waves.................................................................................59

    Fig. 3.9. Spatial distribution of (a) bottom shear stress, (b) dissipation rate averaged in theBBL, (c) entrainment rate at the top of the BBL, and (d) mass transfer coefficient at thesediment-water interface. The values shown are 90th percentile for panel (a) and

    average over the field campaign for panels (b-d). The BBL thickness is assumed to be afunction of depth (Fig. 3.3d), and Sc = 500 (corresponding to O2) is used to estimatethe mass transfer coefficient. Estimations were made only below the thermocline forpanel (b-d). The entrainment rate may be overestimated near the perimeter because thestrong stratification leads to lower mixing efficiency (Lemckert et al. 2004). ........................ 62

    Fig. 3.10. Variables shown in Fig. 3.9 expressed as a function of bottom depth. (a) Bottomshear stress, (b) dissipation rate averaged in the BBL, (c) entrainment rate at the top ofthe BBL, and (d) mass transfer coefficient at the sediment-water interface. Open circlesshow 1-m (arithmetic) bin average of 90th percentile for panel (a) and average over thefield campaign for panels (b-d). The triangles in panel (b) indicate results of directturbulent measurements in the BBL at different locations by Lemckert et al. (2004).The entrainment rate may be overestimated in the metalimnion (shaded area in panel

  • 8/2/2019 Thesis Kenji Shimizu

    11/143

    xi

    [c]) because the strong stratification leads to lower mixing efficiency (Lemckert et al.2004).........................................................................................................................................63

    Fig. 3.11. Absolute value of temporally averaged frequency response function defined as)()(

    1)(

    1)( ~~,

    ~ rrs

    rr eavH

    = . (see Eq. 3.26). Normalized forcing frequency and damping

    rate are defined asf

    r

    f 1)( +

    = and )(1)()( rrr +

    = . Dotted lines, respectively, indicate

    the angular frequencies of V1H1A, V1H3C, V2H1A, V1H1C, and V2H3C from left toright. .........................................................................................................................................65

    Fig. 4.1. Ekman velocities, (uE, vE), when (uI,vI) = (1, 0) and t= 0 (see Eq. 4.14). Profiles (a)to (i) correspond to S = *1f = 0.30, 0.70, 0.90, 0.99, 1.00, 1.01, 1.10, and 3.00

    with 1.5 offset for each profile. Profiles (a) and (i) are almost identical to the Ekmanlayer and the Stokes layer, respectively. Note the sharp transition of the structure near

    S = 1.00. ................................................................................................................................79

    Fig. 4.2. Ekman transport in the rotating bottom boundary layer, (Ebxq , Ebyq ), and sidewall

    boundary layer,Ewyq , when (uI,vI) = (1, 0). (a) Real part (in phase with the far field

    flow), and (b) imaginary part (quarter period ahead of the far field flow). In panel (b),solid and dashed lines coincide for 1

  • 8/2/2019 Thesis Kenji Shimizu

    12/143

    xii

    Fig. 4.6. Dissipative modal structure of Kelvin waves with S = 0.3 (first row), S = 0.6 (secondrow), S = 0.8 (third row), and Poincar waves with S = 0.8 (fourth row) and S = 10(fifth row) in linearly stratified basin withN= 1 and 22SB = . Shading and vector

    show isopycnal displacements and horizontal velocities atz = -0.75, respectively. Thefirst column shows the dissipative modal structure when (E/S)1/2 = 0.05, and the secondshows the inviscid modal structure. The third to fifth columns correspond to first order

    corrections due to the internal wave cancelling due to bottom boundary layer, spin-down, and internal wave cancelling due to sidewall boundary layer, respectively. Thevariables are divided byA in the fifth column, and by 5 and 2 in panels (n) and (t),respectively, for plotting purposes. See caption of Fig. 4.4 for number of modes usedfor plotting. ............................................................................................................................ 100

    Fig. 4.7. Lake Kinneret. (a) Bathymetry, (b) typical stratification at T4 and associatedbuoyancy frequency in summer, and (c) vertical modes under stratification in panel (b).In panel (a), the origin of the figure is situated at 32.70N, 35.51E, and contour linesare drawn every 5 m...............................................................................................................103

  • 8/2/2019 Thesis Kenji Shimizu

    13/143

    xiii

    List of symbols

    Variables

    Roman (lowercase)

    a0 Typical isopycnal displacement (m))(~ ra Modal amplitude ofrth ([quasi-] three-dimensional) mode (-)

    ),(~ srb Modal expansion coefficients (-)

    c Celerity (m s-1)cb, csw Linear friction coefficient (m

    s-1))(~ re Normalizing factor ofr

    th (spatial) mode (J)

    f Coriolis parameter (rad s-1))(~ rf Modal force for r

    th mode (s-1)

    f External force vectorg Accerelation due to gravity (m s-2)h Layer thickness (m)

    i = 1 Imaginary unitk Layer index (-)k

    Vertical unit vector (-)l Length element along horizontal boundary (m)

    ( )Tyx nnn ,= Unit normal vector to boundary (-)

    p Pressure (N m-2)),('~ mlp Dynamic pressure induced by l

    th vertical, mth horizontal three-

    dimensional mode (N m-2)

    Eq Ekman transport (m2 s-1)

    r Radial coordinate (m)

    ( )Tyx sss ,= Horizontal unit normal vector to sidewall (-)

    t Time (s)u x component of velocity (m s-1)

    ),( mlu

    mth horizontal modal velocity that belongs to lth vertical mode

    (m s-1)),(~ mlu Horizontal velocity in induced by lth vertical m

    th horizontal

    three-dimensional mode (m s-1)uE x component of Ekman velocity / Sidewall Ekman normal

    velocity (m s-1)v y component of velocity (m s-1)vE y component of Ekman velocity (m s

    -1)),( mlv

    m

    th horizontal modal velocity that belongs to lth vertical mode

    (m s-1)),(~ mlv Horizontal velocity in induced by l

    th vertical mth horizontal

    three-dimensional mode (m s-1)

    bfv 1

    0

    = Friction velocity (m s-1)

    v Horizontal velocity vector (m s-1) (Chapter 2, 3) / Three-

    dimensional velocity vector (m s-1) (Chapter 4))(r

    kv Velocity vector in kth layer induced by rth mode (m s-1)

    )(~ rkv

    Modal velocity vector in kth layer induced by rth mode (m s-1)

    we Entrainment rate (m s-1)

    wE Ekman normal velocity (m s-1)

    x Horizontal Cartesian coordinate (m)

  • 8/2/2019 Thesis Kenji Shimizu

    14/143

    xiv

    x Horizontal Cartesian coordinate vector (m) (Chapter 2, 3) /

    Three-dimensional Cartesian coordinate vector (m) (Chapter 4)y Horizontal Cartesian coordinate (m)z Vertical coordinate (m)

    Roman (uppercase)

    A Plane area (m2) (Chapter 2, 3) / Aspect ratio (-) (Chapter 4)

    B = f-2N2A2 Stratification parameter (-)C Characteristic celerity (m s-1)C Linear damping operator of shallow water systemCb Bottom drag coefficient (-)D Diffusion coefficient (m2 s-1)

    ][rD Rate of energy dissipation due to +randrmode (J s-1)

    E Total energy of the system (J) (Chapter 2, 3) / Ekmannumber ( )21 = Hf (-) (Chapter 4)

    E(r) Modal total energy ofrth mode (J)

    E[r] Modal total energy of +randrmode (J)

    F Energy flux (J m-2 s-1)

    H Total depth (m)K Linear operator of non-dissipative shallow water systemL Horizontal length scale of a basin / length of a rectangular basin

    ( ) 4/12/2 DLB = Batchelor length (m)

    M Weight matrix

    zgN =10

    Buoyancy frequency (rad s-1)

    O Zero matrixR Radius of a circular basin (m)S = c (fL)-1 Burger number (-)Sc = D-1 Schmidt number (-)T Period (s)

    U Stretched horizontal velocity (m

    s

    -1

    )V Volume (m3)W Stretched vertical velocity (ms-1)X Stretched horizontal coordinate (m)Z Stretched vertical coordinate (m)

    )(rW Complex rate of work done to rth mode (J s-1)

    ][rW Rate of work done to +randrmode (J)

    Greek (lowercase)

    Mass transfer coefficient (m s-1) (Chapter 3) / Vertical wave

    number of bottom boundary layer flow (m-1) (Chapter 4))(r Damping rate ofr

    th mode (rad s-1)

    Thickness of sublayer (m)

    =

    =otherwise

    jiifji 0

    1,

    Kronecker delta

    ( )x

    Dirac delta function

    Error or residual (various) Dissipation rate of turbulent kinetic energy (m2s-3)

    ( )

    ( ) ( )

    >

    ==

    1/

    1/

    01

    01

    k

    k

    kk

    k

    Non-dimensional density difference across k

    th interface (-)

    1= cfR Rossby radius of deformation (m)

    von Karman constant (= 0.41) (Chapter 3)

    Kinematic viscosity (m2

    s-1

    )

  • 8/2/2019 Thesis Kenji Shimizu

    15/143

    xv

    )( l Vertical modal structure (of pressure and horizontal velocities)

    in a continuously stratified basin (-) Volume stream function (m

    2 s-1)

    Azimuthal coordinate (rad)(r) Modal phase ofrth mode (rad) Density (kg m

    -3)

    ( )Tyx ,=

    Shear stress vector (N m-2)

    Angular frequency (rad s-1) State vector of motion for shallow water system

    )(~ r Spatial modal structure ofrth ([quasi-] three-dimensional) mode

    Surface/interface displacement (m))(r

    k Displacement in kth interface induced by rth mode (m)

    )(~ rk Modal displacement in kth interface induced by r

    th mode (m)),( ml

    mth horizontal modal displacement that belongs to lth vertical

    mode (m)

    Greek (uppercase)

    mix Mixing efficiency (-)( )Tf 2/00=

    Local vertical angular velocity vector of Earth (rad s-1)

    Subscripts0 Nominal value / 0th order solution (Chapter 4)1 1st order solution (Chapter 4)b Bottome Vertical variation at the equilibriumf Forcingiwc Internal wave cancelling

    k(>0) Layer indexm Moleculars Surfacesp Spin-downsw Sidewallx x component in Cartesian coordinatey y component in Cartesian coordinatez Vertical componentBBL Bottom boundary layerDSL Diffusive sublayerE Ekman layer / Ekman normal velocityI Far-fieldSW Sediment-water interfaceVSL Viscous sublayer

    Superscriptsg Geostrophic mode(l) Vertical modal index(m) Radial modal index(n) Azimuthal modal index(r) Modal index for (quasi-) three-dimensional mode(p), (q), (s) Dummy modal indices[r] Sum of +rand -rmodew Wave modeL Left modeR Right mode

  • 8/2/2019 Thesis Kenji Shimizu

    16/143

    xvi

    + Non-dimensionalized variable

    Mathematical operators

    ( )diag Diagonal matrix (diagonal components are shown in the

    argument)( )* Complex conjugate

    ( )T Transpose

    ( )H Conjugate transpose (or Hermitian)

    ( ) Temporal average

    ,,,,, rzyxt Partial derivative with respect to the variable in the subscript

    (various) Horizontal differential operator (m-1) (Chapter 2, 3) / Three-

    dimensional differential operator (m-1) (Chapter 4)( ) = dA

    H '', Inner product

  • 8/2/2019 Thesis Kenji Shimizu

    17/143

    xvii

    Acknowledgements

    Looking back the beginning of this journey, it was a fortunate coincidence I came to

    CWR. I wanted to study overseas I wanted to do research that contributes to scientific

    understanding and to write international papers. I found a new Japanese governmentscholarship starting from 2005 few days after the deadline, but there was no one who

    had applied for it and N. Tsutsumida at Tokyo Institute of Technology kindly accepted

    my late application. I knew of CWR as T. Ishikawa, my supervisor in my Masters, had

    long-term collaboration with CWR, and I was accepted by CWR probably because of

    his recommendation.

    I could not have achieved this milestone without the continuous encouragement and

    support of my supervisor, Jrg Imberger. I worked hard as I knew my scholarship wasonly for three years from the beginning, but I would not have been able to finish my

    PhD in three years and a few months without his hard work and patience, particularly

    for last few months. I have achieved what I aimed thanks to Jrg he accepted and

    extended my ideas (I was told originality was not appreciated in Japan!) and helped me

    out writing the ideas down in papers in English. I am happy to have had at least a

    glimpse into world-leading research. I also thank Jrg for giving me opportunities to

    travel to Kenya and South America, where I had never imagined visiting.

    I appreciate collaboration with Dr. Kumagai and Dr. Jiao at Lake Biwa

    Environmental Research Institute, who kindly provided bathymetry, meteorological

    data, and water quality data from Lake Biwa during the first part of my study. They also

    contributed to the study through discussion on formation mechanisms of the gyres.

    I am grateful for valuable and helpful discussions with S. Morillo, P. Okely, P.

    Yeates, A. Gmez-Giraldo, A. de la Fuente, A.M. Simanjuntak, K. Nakayama, T.

    Shintani, and J. Antenucci. I hope I helped them as much as they helped me.

    Particularly, discussion with A. Gmez-Giraldo on preliminary ideas of the modal

    analysis encouraged me in developing the theory, and tough questions from A. de la

    Fuente on the theory contributed making it clearer. P. Okely helped me a lot by

    correcting my English in the early stage of my writing the first paper (how many articles

    she corrected before Jrg read the draft!). I thank J. Antenucci, C.J. Dallimore, A.

    Gmez-Giraldo, T. Johnson, P. Okely, T. Shintani, and anonymous reviewers of the

    first two papers for reading the manuscripts of my papers and giving me constructive

  • 8/2/2019 Thesis Kenji Shimizu

    18/143

    xviii

    and critical comments. R. Alexander, I. Hillmer, P. Okely, and J. Petruniak gave me

    valuable comments on my final thesis presentation.

    Support of staff and friendship of students in CWR and SESE made my study really

    enjoyable. Playing sports and having fun together plus a bit of (a lot of?) distraction

    from my work were great help to me, as I tend to work far too much when I push

    myself. I particularly thank A.M. Simanjuntak for helping me out to settle in CWR and

    Perth and my officemates P. Yeates, S. Morillo, A. de la Funte, C. Boon, P. Okely,

    and P. Huang for being tolerant for me talking to myself loudly while I was

    programming and solving equations! (Actually, I dont have to thank A. de la Funte as

    everyone agrees that he was much noisier than me haha!)

    Finally, to my friends, housemates, and family thank you very much for supportingme during this long tough journey!

    I appreciate financial support of Japanese Government (MEXT) scholarship, Tokyo-

    Tech Long Term Overseas Study Support Program, and ad-hoc CWR scholarship.

  • 8/2/2019 Thesis Kenji Shimizu

    19/143

    xix

    Preface

    This work was completed during the course of my enrolment for the degree of

    Doctor of Philosophy at the Centre for Water Research (CWR), The University of

    Western Australia. The main body of this thesis (Chapter 2 to Chapter 4) is acompilation of three papers written for journal publication. Each chapter is a stand-

    alone manuscript, which includes abstract, literature review, methods, results,

    discussion, and conclusions. The introductory Chapter 1 presents the motivation for this

    study and links the following three chapters. The major outcomes of this work are

    summarized in Chapter 5 followed by recommendations for future work.

    Chapter 2 has been published inLimnology and Oceanography as Shimizu, K., J.

    Imberger, and M. Kumagai. 2007. Horizontal structure and excitation of primarymotions in a strongly stratified lake. Limnol. Oceanogr. 52: 2641-2655. The processing

    of field data and three-dimensional modeling was conducted by myself under

    supervision of Jrg Imberger, and the theory of modal analysis and numerical scheme

    for modal computation were developed by myself. Jrg Imberger also checked scientific

    integrity of the research and edited the manuscript thoroughly, which was originally

    written by myself. Michio Kumagai contributed to the work through provision of field

    data and discussion on dynamics and formation mechanisms of gyres.

    Chapter 3 is in press byLimnology and Oceanography as Shimizu, K., and J.

    Imberger. Energtics and damping of internal waves in a strongly stratified lake. Jrg

    Imberger suggested the original idea of extracting internal wave modes from thermistor

    chain data, and I developed a method to fit numerically computed modes based on the

    modal amplitude equations with estimating the damping rates. All the data processing,

    modal calculation, and mode fitting were done by myself. The manuscript was

    originally written by myself and edited by Jrg Imberger.

    Chapter 4 is to be submitted toJournal of Fluid Mechanics as Shimizu, K., and J.

    Imberger. Damping mechanisms of internal waves in a continuously stratified rotating

    basin. Jrg Imberger identified the problem of how gravity waves are damped by thin

    boundary layer without momentum diffusion during the work presented in Chapter 3.

    He then suggested solving the problem using the perturbation method. I obtained the

    understandings of the damping mechanism referred to as internal wave canceling by

    following previous studies, and further combined it with the spin-down process to

  • 8/2/2019 Thesis Kenji Shimizu

    20/143

    xx

    extend the analyses to a stratified rotating basin. All the mathematical derivations and

    writing were originally done by myself and checked by Jrg Imberger.

    I have the permission of all the co-authors to include the above manuscripts in my

    thesis.

    Kenji Shimizu Jrg Imberger

    (Coordinating supervisor)

  • 8/2/2019 Thesis Kenji Shimizu

    21/143

    Chapter 1. Introduction

    1

    Chapter 1.Introduction

    1.1 MotivationModal analysis has been a useful standard tool in vibration problems of particles and

    solid bodies. These systems have natural periods of oscillation and the associated

    distinct spatial structure inherent to the system, called modes, satisfy general orthogonal

    relationships. Use of modes and their orthogonality reduces the original coupled

    problem of multi-degree of freedom (or partial differential equations) into a set of

    decoupled modal amplitude equations for individual modes that are equivalent to the

    governing equation for a single harmonic oscillator (e.g., Timoshenko et al. 1974). The

    method is applicable even for large number of particles or solid bodies with complicated

    shape, where analytic solutions are difficult to obtain, as the theory is applicable to

    numerically computed modes. Spatial and temporal effects of forcing are also separated

    in this method: spatial correlation between forcing and modal structure determines the

    effectiveness of the forcing exciting the mode, expressed as a modal force term,

    whereas the modal amplitude equations (with the modal force term) describe temporal

    interactions between each mode and the forcing, including excitation, canceling

    (Mortimer 1953), phase shift (Raudsepp et al. 2003; Gmez-Gilaldo et al. 2006), andresonance (Antenucci and Imberger 2003; Gmez-Giraldo et al. 2006), all of which

    have been observed in stratified lakes and semi-enclosed seas. By separating total

    motion into modal components and effects of forcing into spatial and temporal factors,

    modal analysis provides conceptual and general understandings of linear dynamics of

    such systems that are not easily obtainable from experiments or numerical simulations.

    It is well known that this method can be applied to fluid motions in non-rotating

    semi-enclosed shallow basins (e.g., Lamb 1932), such as small lakes. Application of

    linear modal analysis is often beneficial as fluid motions in a geophysical scale are often

    well described by linear theory (e.g., Csanady 1975). Although less attention has been

    paid recently, the method can be extended to stratified rotating basins. Orthogonality of

    the modes was shown by Proudman (1929) for homogeneous rotating shallow basins,

    and Platzman (1972, 1975, 1984) applied modal analysis to numerically compute tides

    in the world ocean. Numerical schemes to calculate modes in stratified basins with

  • 8/2/2019 Thesis Kenji Shimizu

    22/143

    Application of modal analysis to strongly stratified lakes

    2

    arbitrary shape have been proposed, for example, by Schwab (1976) and Buerle

    (1985). However, the theory has not been well developed for stratified rotating basins.

    This thesis is aimed to obtain better and unified understandings of basin-scale

    motions in stratified rotating lakes using modal analysis. This purpose was achieved by

    extending the theory to both layer- and continuously stratified rotating basins with

    arbitrary shape. One of the advantages is that the method provides general and relatively

    simple framework to understand dynamics of basin-scale motions in real lakes through

    numerical computation of modes, unlike analytical studies where the basin shape needs

    to be simplified. Another advantage is the orthogonality of modes that enables us to use

    generalized Fourier series for theoretical study and to extract modal components from

    simulation results and field data, providing new tools for theoretical, numerical, and

    field studies. These points will be illustrated in the following chapters.

    1.2 OverviewFollowing this brief introduction are three chapters where theory of modal analysis

    for both layer- and continuously stratified lakes are developed and applied to better

    understand evolution, excitation, energetics, and damping of basin-scale motions in

    stratified rotating lakes, and the implications on mass transport processes.Excitation of basin-scale motions was investigated in Chapter 2 byextending the

    theory of modal analysis to layer-stratified basins with arbitrary bathymetry. The

    shallow water equations were reduced to modal amplitude equations that describe

    evolution of individual modes forced by wind stresses. It was shown that winds excited

    modes that had similar horizontal velocity structure in the surface layer compared to the

    wind stress pattern. Numerically calculated modes were then used to show how winds

    preferentially excited certain modes in basins with irregular shape. Evolution of modeswas extracted from results of three-dimensional hydrodynamic simulation results in

    order to assess applicability of the linear modal analysis to basin-scale motions in lakes.

    Damping was neglected in this chapter.

    In Chapter 3, the theory was extended to include linear damping and then applied to

    estimate energetics and damping rates of basin-scale internal waves from thermistor

    chain data. The results indicated that basin-scale internal waves were damped within a

    few periods. Bottom friction was considered as a primary cause of the fast damping,

  • 8/2/2019 Thesis Kenji Shimizu

    23/143

    Chapter 1. Introduction

    3

    although it was not clear how bottom friction confined within a thin boundary layer was

    able to damp internal waves quickly. The estimated near-bottom current velocities were

    also used to estimate spatial variability of near-bottom mass transfer processes.

    To answer the question about fast damping of basin-scale internal waves, the

    damping processes were analytically investigated in Chapter 4. The damping

    mechanism was understood as a combination of two fast damping mechanisms: waves

    generated by oscillatory boundary layers cancelling the parent wave (Johns 1968; Mei

    and Liu 1973) and the well-known spin-down (Greenspan 1968; Pedlosky 1979; Gill

    1982) modified by the periodicity. The theory of modal analysis was extended to a

    continuously stratified basin, and applied to flat-bottomed rectangular and circular

    basins in this chapter.

    Major conclusions obtained in these studies and recommendations for future work

    are summarized in Chapter 5.

  • 8/2/2019 Thesis Kenji Shimizu

    24/143

    Application of modal analysis to strongly stratified lakes

    4

    Chapter 2. Horizontal structure and excitation of primary motionsin a strongly stratified lakes

    Abstract

    A modal analysis in the horizontal plane was extended to a layer-stratified basin

    with irregular bathymetry, and the theory was applied to Lake Biwa to investigate the

    horizontal structure and excitation of the basin-scale internal waves and gyres. The

    horizontal structure of the basin-scale internal waves consisted of cyclonic and

    anticyclonic elliptic cells, each of which appeared to follow the dispersion relationship

    of Kelvin and Poincar waves in elliptic basins. The internal waves were preferentially

    excited depending on the arrangement of the cells and the wind direction, but the

    spatial distribution of wind stress curl over the lake primarily determined the horizontalstructure of the ensuing gyres. Decoupled evolutionary equations for the individual

    modes provided a good approximation for excitation of the internal waves and early

    stages of excitation of the gyres before non-linear effects and damping become

    significant. The modal decomposition of hydrodynamic simulation results also showed

    that the primary action of the wind was to excite the internal waves; however, these

    internal waves were damped within a few days and the dynamics during calm periods

    were dominated by the gyres, illustrating the importance of internal waves on mixing

    and gyres on long-term horizontal transport.

    2.1 IntroductionMotions in a lake are primarily energized by surface wind stresses. Wind excites

    primary motions including seiches, basin-scale internal waves and gyres, which in turn

    cause secondary motions, such as high-frequency internal waves and residual

    circulation (Imberger 1998; Okely and Imberger 2007). The spatial structure and

    amplitude of these motions have important implications for chemical and biologicalprocesses in lakes since they determine the flux path of biogeochemicalsubstances

    (e.g., Nishri et al. 2000; Eckert et al. 2002). Understanding of the spatial structure and

    excitation of primary motions provides the foundations for the understanding of

    subsequent physical, chemical, and biological processes.

    Published as: Shimizu, K., J. Imberger, and M. Kumagai. 2007. Horizontal structure and excitation ofprimary motions in a strongly stratified lake. Limnol. Oceanogr. 52: 2641-2655. Centre for WaterResearch Reference ED 2116-KS.

  • 8/2/2019 Thesis Kenji Shimizu

    25/143

    Chapter 2. Horizontal structure and excitation of primary motions

    5

    The horizontal structure of basin-scale internal waves may conveniently be

    visualized in terms of fundamental mode Kelvin and Poincar waves in flat-bottomed

    circular or elliptic basins (Antenucci and Imberger 2001). A Kelvin wave has a

    subinertial frequency and propagates cyclonically around the basin (counterclockwise in

    Northern Hemisphere and clockwise in Southern Hemisphere), and it has a horizontal

    structure that depends on the Burger number ( ) 1= fLcS , where c is the celerity, f the

    Coriolis parameter, and L the horizontal length scale (Antenucci and Imberger 2001).

    When the Burger number is small, the interface displacement and velocity are largest at

    the boundary and decay exponentially offshore with a length scale given by the Rossby

    radius of deformation 1= cfR (Antenucci and Imberger 2001). When the Burger

    number is large, the frequency becomes superinertial, and the horizontal structure

    approaches that of a Poincar wave (Antenucci and Imberger 2001). (Although the

    name, Kelvin wave, is not assigned to superinertial cyclonic waves, we shall use the

    name in this paper for the fundamental mode cyclonic wave.) A Poincar wave

    propagates anticyclonically (clockwise in Northern Hemisphere) with a maximum

    velocity in the middle of the basin and zero velocity at the boundary (Antenucci et al.

    2000).

    In real basins, irregular bathymetry modifies the horizontal structure of a Poincar

    wave by localizing the wave (Wang et al. 2000) or by introducing an accompanying

    cyclonic cell (Gmez-Giraldo et al. 2006); whereas the horizontal structure of a Kelvin

    wave appears insensitive to basin irregularities (Mortimer 1974; Gmez-Giraldo et al.

    2006). Recently Gmez-Giraldo et al. (2006) analysed an internal wave consisting of an

    anticyclonic cell and a cyclonic cell in Lake Kinneret and suggested that the cells were

    governed by the dispersion relationship of Poincar and Kelvin waves, respectively, in a

    circular (or elliptic) basin.

    Excitation of basin-scale internal waves depends on both the temporal and spatial

    variations of the wind forcing. Analytical solutions of internal waves excited by a

    suddenly imposed wind show that the amplitudes increase in the first half of the wave

    period, reach a maximum after half the period and then decrease in the second half

    period (e.g., Birchfield 1969; Stocker and Imberger 2003). Field and modelling studies

    indicate that excitation or cancellation of an internal wave depends on the phase

    between the wave and the wind (Mortimer 1953;Antenucci et al. 2000; Rueda et al.

  • 8/2/2019 Thesis Kenji Shimizu

    26/143

    Application of modal analysis to strongly stratified lakes

    6

    2003), and that winds not only excite internal waves but also shift their observed phase

    due to generation of additional internal waves (Gmez-Giraldo et al. 2006). The

    effectiveness of winds in excitation or cancellation also depends on the spatial

    distribution of the wind stress field over the lake and the horizontal structure of internal

    waves; however, this interplay is not well understood in an irregular basin.

    Unlike basin-scale internal waves that are excited directly by wind, gyres may be

    excited by wind stress curl (Emery and Csanady 1973; Endoh 1986), topographic

    effects (Csanady 1973), residual current resulting from internal waves (Ou and Bennett

    1979), and thermal effects (Huang 1971; Schwab et al. 1995). Although thermal effects

    can be important on a seasonal timescale (Schwab and Beletsky 2003; Akitomo et al.

    2004), recent studies indicate that, for timescales shorter than seasonal, wind stress curl

    and topographic effects are the major driving forces of gyres (Laval et al. 2003, 2005;

    Rueda et al. 2005) and that the contributions from internal waves are minor (Pan et al.,

    2002). We will limit our focus to wind-driven gyres in this paper.

    The spatial structure of wind-driven gyres depends strongly on the spatial

    distribution of the wind (Pan et al. 2002) and bathymetry(Csanady 1973). Theoretical

    analysis of flat basins without the variation of the Coriolis parameter (the f -plane)

    indicates that the horizontal structure of gyres is determined by external forcing(Proudman 1929; Gill 1982). Variations in depth induce a topographic gyre due to

    larger specific momentum input in a shallow coastal region compared to that in deeper

    regions (Csanady 1973). A topographic gyre degenerates primarily into topographic

    waves after the wind ceases (Raudsepp et al. 2003), and propagates cyclonically around

    the basin with the periods and horizontal structure determined primarily by the basin

    shape and bottom slope (Rhines 1969). In strongly stratified lakes, currents associated

    with gyres are observed only in the epilimnion, and the geostrophic pressure associated

    with the circulation induces doming (or depression) of the pycnocline (Endoh et al.

    1995a; Kumagai et al. 1998; Laval et al. 2005).

    Lake Biwa is the largest lake in Japan, consisting of the relatively deep North Basin

    and shallower South Basin. This lake will be used to illustrate the theoretical

    developments made in this paper, as several basin-scale internal waves and a gyre

    system consisting of two or three gyres have previously been identified. The dominant

    internal wave is the vertical and horizontal mode 1 Kelvin wave (Kanari 1975), and

  • 8/2/2019 Thesis Kenji Shimizu

    27/143

    Chapter 2. Horizontal structure and excitation of primary motions

    7

    second, third, and fourth horizontal mode basin-scale internal waves have also been

    observed after strong winds (Saggio and Imberger 1998). The largest and most

    persistent cyclonic gyre in the north of the North Basin (First Gyre) is often

    accompanied by an anticyclonic gyre in the middle of the basin (Second Gyre) both of

    which induce circular currents with a characteristic water velocity of 0.1 m s-1 (Endoh

    and Okumura 1993). An unstable cyclonic gyre has also been found in the south of the

    North Basin (Third Gyre), and long-term current measurements have shown that the

    locations and the number of gyres change on a seasonal timescale (Kumagai et al.

    1998). It has not, however, been shown why the internal waves of higher horizontal

    modes are excited and what determines the horizontal structure of the gyres.

    The purpose of this paper is to obtain a better and more comprehensive

    understanding of the spatial structure and excitation of internal waves and gyres by

    using a modal analysis in the horizontal plane. This technique has been previously used

    to analyse the horizontal structure of tides and seiches in homogeneous basins

    (Platzman 1972; Rao andSchwab 1976) and internal waves in stratified basins (Buerle

    1985, Lemmin et al. 2005). However, the analysis also enables us to derive a set of

    decoupled evolutionary equations for the individual modes (e.g., Proudman 1929;

    Lighthill 1969;Platzman 1984) that describe the effects of winds on the primary

    motions. Simulations with the three-dimensional hydrodynamic model ELCOM

    (Estuary, Lake and Coastal Ocean Model, Hodges et al. 2000; Laval et al. 2003;

    Simanjuntak et al. Submitted) are also used to confirm the results obtained from the

    semi-analytical modal analysis.

    This paper is structured as follows. First, the field data used in this paper will be

    briefly described.Second, we will describe the extension of the theory of modal

    analysis in the horizontal plane to a layer-stratified basin and demonstrate its application

    to Lake Biwa. Third, we will illustrate how the modal components can be extracted

    from the results of the three-dimensional hydrodynamic simulations to confirm the

    results of the modal analysis. Finally, we will discuss some implications of the results

    for a strongly stratified lake in general.

  • 8/2/2019 Thesis Kenji Shimizu

    28/143

    Application of modal analysis to strongly stratified lakes

    8

    2.2 Field DataThis study uses field data collected in 1993 and 1994 for wind speed and direction,

    temperature profile, and velocity structure in the water column in Lake Biwa. Wind data

    were measured at five land-based stations around the lake by the AutomatedMeteorological Data Acquisition System (AMeDAS; managed by Japan Meteorological

    Agency) and at BN50 during Biwako Transport Experiment (BITEX93; 21 Aug -16

    Sep 1993) (Fig. 2.1).Since the wind speed measured at land-based stations was

    considerably lower than that measured at BN50, the wind speed measured at the land-

    based stations were correlated to the wind speed at BN50 and then corrected by

    multiplying by a factor of 2.5. Temperature profiles were measured by a thermistor

    chain at BN50 during BITEX93 and as part of biweekly routine measurements at Sta.

    17B in 1994 (Fig. 2.2). The thermistor chains had 20 thermistors spaced every 1 m in

    the thermocline, extending up to 5 m apart near the surface and the bottom where the

    stratification was weaker (seeSaggio and Imberger [1998] for details). Vertical profiles

    of water velocity were measured by a shipboard broad-band acoustic Doppler current

    profiler (ADCP) on a monthly basis in 1994 along 11 transects in W-E direction,

    covering the whole North Basin (seeKumagai et al. [1998] for details).

  • 8/2/2019 Thesis Kenji Shimizu

    29/143

    Chapter 2. Horizontal structure and excitation of primary motions

    9

    N

    36.45 N135.90 E

    Imazu

    Hikone

    Ohtsu

    Torahime

    Minami Komatsu

    Shiozu bay

    BN50

    Land-based meteorological

    Thermistor chain & wind sensor

    * contour line every 10 m

    Sta.17B

    Routine temperature profiling

    a)

    Ab) Plan view of computat ional grid in A

    c) Computational grid along

    B-B (modal analysis)

    d) Schematic computational grid

    along B-B (ELCOM)

    ,huvw

    B B

    S

    outh

    Basin

    Nor

    thB

    asi

    n

    50 10km

    15

    Fig. 2.1. Bathymetry of Lake Biwa and computational grids: (a) bathymetry of Lake Biwa and

    locations of measurement stations, (b) magnified plan view of computational grid in A (common for

    both the modal analysis and ELCOM simulations), (c) computational grid along cross-section B-B

    used for the modal analysis, and (d) schematic computational grid along cross-section B-B used for

    the ELCOM simulations (vertical grid resolution is finer than the schematic). Thin and thick lines

    in panel (a) show bathymetry contour with 10-m interval and horizontal boundaries of

    computational grid at the surface and 18-m deep, respectively. The dashed line indicates a

    boundary between the northern and southern halves that are used for the calculation of spatially

    averaged wind stress and its curl (see Fig. 2.7).

  • 8/2/2019 Thesis Kenji Shimizu

    30/143

    Application of modal analysis to strongly stratified lakes

    10

    5 10 15 20 25 30

    0

    10

    20

    30

    40

    50

    60

    70

    80

    90

    100

    Temperature (C)

    Depth(m)

    25 Aug 199320 Jun 199418 Jul 19942 layer (1993)

    Fig. 2.2. Temperature profiles used for the modal analysis and the hydrodynamic simulations. The

    profiles in 1993 and 1994 were measured at BN50 and Sta. 17B, respectively. For the modal

    analysis with two-layer stratification, the depth of the interface was set to 18 m (corresponds to

    maximum buoyancy frequency).

    2.3 Theory of modal analysis2.3.1 Basic equations and solutions

    The modal analysis used in this paper is based on the linearized shallow waterequations for a layer-stratified systemwith the Boussinesq and hydrostatic

    approximations(e.g., Monismith 1985; Lemmin et al. 2005). For an incompressible

    fluid in a two-layer system, multiplying the equation of conservation of mass for the

    upper (lower) layer by the acceleration due to gravity (reduced gravity) and including

    layer thickness in the equations of motion gives:

    ( ) ( ) ( )tkxtkxitkxt ,,,,,,

    fKM += (2.1)

    where

    ( ) ( )Tvvtkx 2121,,

    = (2.2)

    is the state vector of motion,

    ( ) ( )Tstkx 000,,

    =f (2.3)

    is the external force vector,

  • 8/2/2019 Thesis Kenji Shimizu

    31/143

    Chapter 2. Horizontal structure and excitation of primary motions

    11

    ( )

    =

    22212

    111

    22

    2111

    0

    20

    020

    000

    00

    ,

    hghgh

    hgh

    hg

    hghg

    ikx

    K (2.4)

    is the linear operator of the shallow water system,

    ( )

    =

    2

    1

    2

    1

    0

    000

    000

    000

    000

    ,

    h

    h

    g

    g

    kx

    M (2.5)

    is the weight matrix, t is the time, ( )Tyxx ,= represents the horizontal coordinates, k

    ( 1, 2= ) is the layer index, is the surface or interface displacement, ( )Tvuv ,= are the

    layer-averaged velocities, g is the acceleration due to gravity, h is the layer thickness,

    ( )Tf 2/00=

    is the local vertical angular velocity due to Earths rotation, f is the

    Coriolis parameter, ( )Tsysxs ,= are the surface stresses, is the density, 011 / =

    and ( ) 0122 / = are the non-dimensional density differences across the surface and

    interface, respectively, 1=i is the imaginary unit, t is the temporal differential

    operator, and ( )Tyx = , is the two-dimensional horizontal differential operator.

    Subscripts 1 and 2 denote the upper and lower layers respectively, subscript 0 denotes a

    reference value, superscript Tstands for the transpose and bold face is used to denote

    mathematical vectors and matrices.

    Multiplying Eq. 2.1 by 1M yields an equation with the Laplaces (tidal) operator

    used by Platzman (1972) and Buerle (1985); however, we choose to retain M for

    convenience as shown below. The boundary condition on each layer is given by zerovolume flux normal to the boundary:

    ( ) 0= kkk nvh

    (2.6)

    where n

    is the normal vector to the boundary.

  • 8/2/2019 Thesis Kenji Shimizu

    32/143

    Application of modal analysis to strongly stratified lakes

    12

    Assuming basin-scale coherent motion and no external forcing, separation of

    variables of the form tiekxtkx ),(~

    ),,(

    = converts Eq. 2.1 into a generalized

    eigenvalue problem:

    KM~~

    = (2.7)

    Solutions to Eq. 2.7 include linearly independent solutions of the form:

    ( ) ( )tir

    r

    ekxtkx)(

    ,~

    ,, )(

    = (2.8)

    where ( )Trrrrr vv )(2)(1)(2)(1)(~~~~~ = are the spatial modal structureand )(r are the

    natural angular frequencies, and ( )r in the superscript are the modal indexes (note that

    the absolute value of )(~ r are arbitrary since they are eigenfunctions; a tilde denotes

    variables whose magnitude depends on the normalization factor ( )(~ re in Eq. 2.15)

    throughout this paper). Corresponding to every solution in Eq. 2.8, there is a conjugate

    solution whose spatial modal structure and natural angular frequency are given by*)(~ r

    and )(r , where * stands for the complex conjugate.

    Solutions ofEq. 2.7, with the boundary condition Eq. 2.6, consist of four classes of

    modes: Gravity modes, vorticity modes, geostrophic modes, and quiescent modes

    (Platzman 1975; Rao and Schwab 1976). The first three modes correspond to primary

    motions, while quiescent modes are associated with spatially uniform change of the

    surface or interface level and induce no motion. Gravity modes consist of surface waves

    (seiches) and internal waves, both being in a balance (to the first approximation)

    between gravity and inertia, and vorticity modes include planetary Rossby waves and

    topographic waves, where the Coriolis force balances inertia. Geostrophic modesrepresent steady circulation that are in geostrophic balance and evolve under external

    forcing. Although geostrophic modes do not appear in a rotating homogeneous basin

    with variable depth (Platzman 1975), they are supported in a stratified basin if the

    variation of the Coriolis parameter is neglected, as may be shown as follows.

    Assuming steady state and no forcing, we may take the curl of the equations of

    motion Eq. 2.1, which yields:

  • 8/2/2019 Thesis Kenji Shimizu

    33/143

    Chapter 2. Horizontal structure and excitation of primary motions

    13

    ( ) ( ) 0// 11111 == vhffghf

    (2.9)

    and

    ( ) ( ) ( ) 0// 2222112 ==+ vhffgghf

    (2.10)

    where the geostrophic balance between the pressure gradient and Coriolis force is used

    to derive the second expression. These equations form the conditions for the existence

    of geostrophic modes. If the variation of the Coriolis parameter is neglected, the layer

    thickness must be constant for any surface or interface displacement (hence velocity) to

    retain a degree of freedom. Although this condition is not satisfied in the lower layer

    and the coastal zone of the upper layer where the water column is not stratified,

    geostrophic modes do exist in the upper layer of a lakes interior where 01 =h . If thevariation of the Coriolis parameter (the -effect) is included, a gradient of potential

    vorticity converts such geostrophic modes into Rossby waves.

    2.3.2 Some property of modesThe modes discussed above have some useful properties. To see this, consider the

    inner product:

    , ' 'H dA= (2.11)

    where the superscript H stands for the conjugate transpose, the prime denotes different

    state of motion and the integral is taken over the whole domain. As shown in Appendix

    2.A, the operator K , along with the boundary condition (Eq. 2.6), is Hermitian (or self-

    adjoint):

    ', ',= K K (2.12)

    This property leads to the orthogonality of modes unless modes are degenerate, or more

    than one mode has the same natural angular frequency (Proudman 1929),

    ( ) ( ) ( ),,

    r s r

    r se = M (2.13)

    ( ) ( ) ( ) ( ),,

    r s r r

    r se = K (2.14)

  • 8/2/2019 Thesis Kenji Shimizu

    34/143

    Application of modal analysis to strongly stratified lakes

    14

    wheresr, is the Kronecker deltaand

    )(~ re is twice the total energy of the rth mode

    when the amplitude of the mode is unity:

    ( )( ) ( ) ( ) ( ) ( ) ( ) ( ) ( ) ( )0 1 1 1 2 2 2 1 1 1 2 2 2, , , ,r r r r r r r r r e g g v h v v h v = + + + (2.15)

    Another useful property is the completeness. Any analytic state vector may be

    expressed in a convergent series of the form

    ( ) ( ) ( )+

    =

    =r

    rr takxtkx)()( ~,

    ~,,

    (2.16)

    where )(~ ra is the complex modal amplitude of rth mode that can be calculated from:

    (Proudman 1929)

    ( ) ( ) ( )( ) ( ) 1 ( ) , , , ,r r ra t e x k x k t = M (2.17)

    The absolute value )(~ ra and angle ( )r , defined as

    ( )( )( )( )( )

    ( )

    ( ) 1

    ( )

    Imtan

    Re

    r

    r

    r

    a tt

    a t

    =

    (2.18)

    determine the amplitude and phase of the mode, respectively. The completeness means

    that the series expansion (Eq. 2.16) is convergent in the weighted square mean sense:

    ( ) ( )( )1

    2r

    r

    E t E t +

    =

    = = ,M (2.19)

    where E is the total energy of the system, and

    ( ) ( )

    2)()()( ~~2

    1taetE

    rrr

    = (2.20)

    the modal total energy of rth mode.

    2.3.3 Decoupled dynamic equations of individual modesThe orthogonality of modes enables us to derive a set of decoupled evolutionary

    equations for each mode as was shown by Lighthill (1969) and Timoshenko et al.

    (1974). For our case, this may be achieved by substituting the series expansion (Eq.

  • 8/2/2019 Thesis Kenji Shimizu

    35/143

    Chapter 2. Horizontal structure and excitation of primary motions

    15

    2.16) to Eq. 2.1, changing the modal index from r to s , taking the inner product with a

    particular mode )(~ r , and using the orthogonality (Eqs. 2.13 and 2.14). This yields:

    ( ) ( ) ( )tftaeitae rrrrrtr )()()()()()( ~~~~~

    += (2.21)

    where the modal force )(~ rf is defined by:

    ( ) ( ) ( ) ( ) ( )( ) ( ) ( )1, , , , , ,r r r

    sf t x k x k t v x x t = = f

    (2.22)

    Since )(r is constant and )(~ ra and )(~ rf are functions of time only, Eq. 2.21is a first

    order ordinary differential equation, and the solution may easily be found. In order to

    show the effects of wind forcing on primary motions, let us further derive the

    evolutionary equations for the modal energy and phase.Writing)(

    )()( ~~ rirr eaa

    = in Eq.

    2.21, taking the temporal derivative on the LHS, multiplying by)(

    )(~ rir ea

    , separating

    real and imaginary parts, and dividing the imaginary part by )(2 rE , we find the

    evolutionary equations of modal energy and phase:

    ( ) ( )( )tWtE rrt)()( Re = (2.23)

    ( ) ( )( ) ( )tEtWt rrrrt)()()()( 2Im += (2.24)

    where )(rW is the complex (in mathematical sense) rate of work done by the wind on the

    rth mode

    ( ) ( ) ( ) ( ) ( )( ) ( )* ( ) ( )1 , , ,r r r r

    sW t a t f t v x t x t = =

    (2.25)

    and )(1 rv

    the average velocity in the upper layer induced by the rth mode. Eqs. 2.20 and

    2.23-2.25provide a general framework for the effects of wind stress on each mode.

    First, note that )(1rv

    is a complex variable whose real and imaginary part that represents

    the velocity field at the present time and a quarter period before the present time,

    respectively (e.g., if the middle panel in Fig. 2.3a-d is assumed to represent the real part,

    then right panel shows -1 times the imaginary part). If )(1rv

    and s

    are parallel

    everywhere, )(rW is real and the wind excites (or cancels) the mode when they are in

  • 8/2/2019 Thesis Kenji Shimizu

    36/143

    Application of modal analysis to strongly stratified lakes

    16

    the same (opposite) direction (Eqs. 2.23 and 2.25). On the other hand, when the upper

    layer velocity and wind stress are normal to each other everywhere, s

    and the

    imaginary part of )(1rv

    are in the same (opposite) direction (hence )(rW is imaginary)

    and the wind accelerates (decelerates) the phase evolution (Eqs. 2.24 and 2.25).

    It is worthy to note that the decoupling shown above is possible in a basin with

    arbitrary bathymetry, once modes have been obtained either analytically or numerically.

    Further, these results obtained for a two-layer basin hold in general, as shown in

    Appendix 2.A.

    2.4 Modal analysis of Lake Biwa2.4.1 Numerical method

    To apply the modal analysis to irregularly shaped Lake Biwa, Eq. 2.7 was discretized

    and solved numerically. A rectangular grid with horizontal spacing of approximately

    460 m 570 m was used with surface and interface displacements defined in the

    middle of each grid and velocity components defined on each face in the direction of the

    velocity (staggered or Arakawa C-grid; see Fig. 2.1b). A finite difference method is

    used for the discretization, and the Coriolis force term was discretized with the method

    suggested by Platzman (1972) to keep the discretized operator K Hermitian, which was

    important in order to retain the orthogonality and completeness of modes in the discrete

    space. Based on the thermistor chain data collected during BITEX93 (Fig. 2.2), the

    thermocline depth and density difference were set to 18 m and 2.27 kg m-3, respectively.

    The phase of all the waves was arbitrarily set to zero when the total volume transport of

    water in upper layer was from north to south.

    Unlike previous studies (Buerle 1985; Lemmin et al. 2005), regions shallower thanthe thermocline were included in the calculation. Wherever the interface intersected the

    bottom, the volume flux normal to the horizontal boundary in the lower layer was set to

    zero (the boundary was assumed locally vertical; see Fig. 2.1c), and continuity of

    displacements andvolume fluxes were applied to the upper layer. Since the method

    assumes infinitely small amplitude, the boundary condition applied to the lower layer

    was essentially same as the method used by Buerle (1985) and Lemmin et al. (2005),

    who assumed a vertical wall along the line where the thermocline intersects the bottom

  • 8/2/2019 Thesis Kenji Shimizu

    37/143

    Chapter 2. Horizontal structure and excitation of primary motions

    17

    and neglected the shallow coastal regions. Note that the inclusion of the shallow regions

    did not affect the self-adjointness of K and properties of the associated modes (see

    Appendix 2.A).

    Table 2.1. Some characteristics of the internal waves and gyres in Lake Biwa. Some of thetopographic waves and surface waves (seiches) are also included to illustrate the spectrum of the

    modes. The surface and internal waves are numbered in decreasing order of their natural periods,

    while the geostrophic gyres are numbered in the opposite order. No number is assigned to

    topographic waves.

    Abbreviations are HM: horizontal mode, PE: percentage of the available potential energy to the

    total energy, [ ]rE : mean modal total energy in the hydrodynamic simulation during BITEX 93, GS:

    geostrophic gyre (calculated as Rossby wave in the upper layer), TW: topographic wave, IW:

    (vertical mode 1) internal wave, SW: surface wave (seiche), CG: cyclonic gyre, AG: anticyclonic

    gyre, CC: cyclonic cell, AC: anticyclonic cell, NL: nodal line, NB: North Basin, N: north, S: south,

    and M: middle.

    Inertial period is 20.7 h. Calculated as the ratio of sum of the first two terms in Eq. 2.15 to

    sum of all the terms. # The sign [r] in the superscript indicates the sum of the conjugate pair, which

    have positive and negative angular velocity with the same magnitude and the same modal structure.

    References in the last column indicate the report of similar structure.

    Data sources are 1) Endoh (1986), 2) Endoh and Okumura (1993), 3) Kumagai et al. (1998), 4)

    Kanari (1975), 5) Saggio and Imberger (1998), 6) Endoh et al. (1995a), 7) Okamoto and Endoh

    (1995), and 8) Kanari (1974). All data except 8) are based on field observation during summer,while 8) is based on a two layer model where the depths of the North and South Basin are taken as

    50 m and 5 m, respectively.

    Class HMPeriod

    this study

    Periodotherstudy

    PE

    (%)

    [ ]rE#

    (GJ)Horizontal structure

    1 6.8 yr - 6.7 2.561 CG in N of NB, corresponding to the FirstGyre1,2,3

    2 9.0 yr - 4.3 0.95 A pair of CG and AG in N of NB3 10.0 yr - 2.5 0.27 3 CGs in N and M of NB, and 1 AG in S of NB

    GS

    4 11.1 yr - 1.8 0.19 2 CGs and 3 AGs

    - 4.97 yr - 0.2 0.04 1 gyre in NW coast of NBTW - 4.70 day -

  • 8/2/2019 Thesis Kenji Shimizu

    38/143

    Application of modal analysis to strongly stratified lakes

    18

    Although the variation of the Coriolis parameter has, in general, a minor effect in

    lakes of the size of Lake Biwa, it was nevertheless included in order to avoid

    degeneracy of geostrophic modes in the numerical solution; the geostrophic gyres thus

    appeared as Rossby waves. However, we prefer to call these modes geostrophic gyres

    in this paper because they correspond to quasi-steady geostrophic circulation in

    practical terms, as their periods (Table 2.1) are much longer than both the inertial period

    and the estimated damping timescale of the gyres (10-20 d) (Endoh 1986).

    2.4.2 Horizontal structure and excitation of internal wavesThe vertical mode 1, horizontal mode 1 (hereafter V1H1) internal wave was a Kelvin

    wave that rotated cyclonically around the basin, the largest interface displacement

    occurred at the northern and southern ends, and the water velocity was nearly parallel tothe thalweg(Fig. 2.3a). The V1H2 and V1H4 internal waves had two and three cells

    where the crests and troughs of the interface rotated cyclonically. The particle orbits

    were nearly parallel to the thalweg except the middle of the basin in V1H4, where the

    velocity vectors appeared to rotate anticyclonically (Fig. 2.3b,c; note that velocity

    vectors of a cyclonic wave of higher horizontal mode do not necessarily rotate in

    cyclonic direction as seen in the analytical solution for a flat-bottomed circular basin;

    e.g., Stocker and Imberger 2003). In V1H2, resonance in Shiozu Bay made the interface

    displacement in the bay larger than in the main part of the lake. V1H3 had a similar

    structure to V1H2 except that the phase of the motion was opposite in Shiozu Bay (not

    shown). V1H6 was the lowest mode with an anticyclonic cell located in the middle of

    the North Basin (Fig. 2.3d). Field data has shown an anticyclonic rotation of the current

    vectors in the middle of the North Basin with a period of 11 h in summer ( Endoh et al.

    1995a), confirming existence of this mode. Overall, the frequencies and horizontal

    structure of these modes matched well with previous studies (Table 2.1).

    The horizontal structure of the velocity in the upper layer determines the potential

    magnitude of the excitation of internal waves under any given wind field (Eq. 2.25; the

    evolution of internal waves and temporal variation of the wind must be taken into

    account to evaluate the actual magnitude as indicated by Eq. 2.23). As an example, we

    considered the effect of the asymmetric shape of the lake and wind directionunder

    spatially uniform winds. A NE-SW wind favoured the V1H1 internal wave (Fig. 2.4)

    because the velocity in the surface layer was approximately parallel to the thalweg (Fig.

  • 8/2/2019 Thesis Kenji Shimizu

    39/143

    Chapter 2. Horizontal structure and excitation of primary motions

    19

    2.3a). On the other hand, WNW-ESE winds would excite V1H2 and V1H4 to a greater

    extent than V1H1 (Fig. 2.4). The velocity pattern associated with V1H2 (Fig. 2.3b)

    showed that the curvature of the thalweg made the WNW-ESE components of the upper

    layer velocity in the two cells parallel and in the same direction, resulting in a large

    value of )(rW . The velocity in the central and southern cells in V1H4 had a similar

    magnitude but opposite direction (Fig. 2.3c), so the correlation between winds and

    upper layer velocity in the northern cell determined the magnitude of excitation.

    Excitation of V1H6 had less dependence on wind direction because the dominant

    anticyclonic cell had a nearly circular particle orbit.

    a) V1H1 b) V1H2

    c) V1H4 d) V1H6

    2 cm s-1

    10 km

    0 0.1 0.2 0.3 0.4 0.5 0.6

    Amplitude (m)

    0 0.2 0.4 0.6

    Interface displacement (m)

    -0.6 -0.4 -0.2

    Fig. 2.3. Internal waves in Lake Biwa: (a) V1H1, (b) V1H2, (c) V1H4, and (d) V1H6. The middle

    panels correspond to the phase where the rates of energy input ( ( ))(Re rW ) from spatially uniform

    winds are the largest ( )(r = 30, 120, 115, 70,see Fig. 2.4), and the right panels correspond to a

    quarter period after the left. Shading and vectors show the interface displacement and the average

    velocity in the upper layer, respectively. In the smaller left panels, shading and lines respectively

    show co-range contour and co-phase lines of wave trough with 45 interval. The co-phase lines with

    triangles correspond to the phase shown in the middle figures, and the triangles show direction of

    rotation. The modes are normalized to )(~ re = 2 GJ. Ellipses with dashed lines indicate cyclonic and

    anticyclonic cells used to calculate the dispersion relationship (see Fig. 2.11).

  • 8/2/2019 Thesis Kenji Shimizu

    40/143

    Application of modal analysis to strongly stratified lakes

    20

    0 90 180 270 3600

    0.1

    0.2

    0.3

    0.4

    0.5

    Wind direction (degrees)

    Normaliz

    edenergyinput(-) V1H1

    V1H2V1H4V1H6

    Fig. 2.4. Normalized maximum energy input from spatially uniform winds blowing from different

    directions. Since energy input depends on the phase of the mode, the maximize energy input is

    normalized by the norm of )(~ r and f (thus vertical axis corresponds to ( ) ( ) ( ) 1/ , ,r r rW M f M f ).

    Wind direction is 0 when the wind is blowing from the north, and it is positive clockwise.

    2.4.3 Spatial structure and excitation of gyresThe horizontal mode 1 geostrophic gyre existing in the upper layer (L1H1) had one

    cell in the middle of the North Basin (Fig. 2.5a), corresponding to the First Gyre. L1H2

    consisted of two counter-rotating gyres, and the superposition with L1H1 led to the

    north-south migration of the First Gyre (Fig. 2.5a,b). L1H3 induced strong gyres in

    south-central part of the North Basin (Fig. 2.5c), corresponding to the Second Gyre.

    General characteristics of these modes, such as quasi-geostrophic circular currents in the

    upper layer and an isostatic balance of the interface displacements, also agreed with the

    field observations (Endoh 1986;Endoh et al. 1995a).

    Geostrophic gyres in the lakes interior are excited only when wind stress curl is

    non-zero. This follows directly from the vorticity equation, but may also be shown from

    Eq. 2.25if we rewrite this equation in terms of the volume stream function for the upper

    layer)(

    1

    r

    :

    )(1

    11

    )(1

    rr hv =

    (2.26)

    that is valid when 1 1f

  • 8/2/2019 Thesis Kenji Shimizu

    41/143

    Chapter 2. Horizontal structure and excitation of primary motions

    21

    Using the stream function and the boundary condition, Eq. 2.25 may be written as

    ( ) ))dAhdAhW srsrr

    ==

    11

    )(1

    )(1

    11

    )(

    (2.28)

    Since the velocity induced by a geostrophic gyre exists in the lakes interior where 1h is

    constant, the above equation shows that geostrophic gyres are excited only if wind stress

    has a non-zero curl component. It also implies that the circulation pattern of geostrophic

    gyres is a reflection of the wind stress curl field, since the horizontal structure of

    geostrophic gyres is rather arbitrary (provided that Eqs. 2.9 and 2.10are satisfied) and

    the correlation between the stream function and wind stress curl determines the

    excitation (Eqs. 2.23 and 2.28).

    a) L1H1 b) L1H2 c) L1H3

    5 cm s-1

    10 km

    -0.3

    0

    0.3

    Interfacedisplacement(m)

    Fig. 2.5. Geostrophic gyres in Lake Biwa: (a) L1H1, (b) L1H2, and (c) L1H3. The panels show the

    most frequently observed phase during days 24