18
1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration of internal waves in lakes with sloping topography L. Boegman, 1 G. N. Ivey, and J. Imberger Centre for Water Research, The University of Western Australia, Crawley, Western Australia 6009, Australia Abstract In a laboratory study, we quantified the temporal energy flux associated with the degeneration of basin-scale internal waves in closed basins. The system is two-layer stratified and subjected to a single forcing event creating available potential energy at time zero. A downscale energy transfer was observed from the wind-forced basin-scale motions to the turbulent motions, where energy was lost due to high-frequency internal wave breaking along sloping topography. Under moderate forcing conditions, steepening of nonlinear basin-scale wave components was found to produce a high-frequency solitary wave packet that contained as much as 20% of the available potential energy introduced by the initial condition. The characteristic lengthscale of a particular solitary wave was less than the characteristic slope length, leading to wave breaking along the sloping boundary. The ratio of the steepening timescale required for the evolution of the solitary waves to the travel time until the waves shoaled controlled their development and degeneration within the domain. The energy loss along the slope, the mixing efficiency, and the breaker type were modeled using appropriate forms of an internal Iribarren number, defined as the ratio of the boundary slope to the wave slope (amplitude/wavelength). This parameter allows generalization to the oceanographic context. Analysis of field data shows the portion of the internal wave spectrum for lakes, between motions at the basin and buoyancy scales, to be composed of progressive waves: both weakly nonlinear waves (sinusoidal profile with frequencies near 10 24 Hz) and strongly nonlinear waves (hyperbolic–secant-squared profile with frequencies near 10 23 Hz). The results suggest that a periodically forced system may sustain a quasi-steady flux of 20% of the potential energy introduced by the surface wind stress to the benthic boundary layer at the depth of the pycnocline. The turbulent benthic boundary layers located at the pe- rimeter of stratified lakes and oceans serve as vital pathways for diapycnal mixing and transport (Ledwell and Hickey 1995; Goudsmit et al. 1997; Wu ¨est et al. 2000). Within the littoral zone, this mixing drives enhanced local nutrient flux- es and bioproductivity (e.g., Sandstrom and Elliott 1984; Os- trovsky et al. 1996; MacIntyre et al. 1999). Indirect obser- vations suggest that turbulent benthic boundary layers are energized by internal wave activity. In lakes, the turbulence occurs as a result of (1) bed shear induced by basin-scale currents (Fischer et al. 1979; Fricker and Nepf 2000; Gloor et al. 2000) and (2) breaking of high-frequency internal waves upon sloping topography at the depth of the metal- imnion (Thorpe et al. 1972; MacIntyre et al. 1999; Michallet and Ivey 1999). While the former process is relatively well understood, the latter remains comparatively unexplored. In lakes where the effects of the Earth’s rotation can be neglected, the internal wave weather may be characterized 1 To whom correspondence should be addressed. Present address: Coastal Engineering Laboratory, Department of Civil Engineering, Queen’s University, Kingston, Ontario K7L 3N6, Canada. Acknowledgments The authors thank K. G. Lamb, B. R. Sutherland, E. J. Hopfinger, J. J. Riley, G. W. Wake, and two anonymous reviewers for com- ments and/or discussions on this work. We also thank Herve ´ Mich- allet for supplying the laboratory data shown in Fig. 10. The Lake Pusiano data were collected by the Field Operations Group at CWR and the Italian Water Research Institute—in particular we thank Diego Copetti. The wavelet software was provided by C. Torrence and G. Compo. L.B. was supported by an International Postgraduate Research Scholarship and a University Postgraduate Award. This research was funded by the Australian Research Council and forms Centre for Water Research reference ED 1665-LB. by the relative magnitudes of the period T i of the horizontal mode-one internal seiche and the characteristic time scale of the surface wind stress, T w (see Stevens and Imberger 1996, their table 1 for typical values). When T w . T i /4, an ap- proach to quasi-equilibrium is achieved, with a general ther- mocline tilt occurring over the length of the basin (Spigel and Imberger 1980; Stevens and Imberger 1996, and Fig. 1b, c). From this initial condition, the resulting internal wave field may be decomposed into a standing seiche, a progres- sive nonlinear surge, and a dispersive high-frequency inter- nal solitary wave (ISW) packet (Boegman et al. 2005). For some lakes, T w K T i , and the progressive surge is believed to originate as a wind-induced wave of depression at the lee shore (see Farmer 1978, and Fig. 1d). This wave of depres- sion will progress in the windward direction, steepening as it travels, and eventually forming a high-frequency ISW wave packet. In the absence of sloping topography, as much as 20% of the available potential energy (PE o )—resulting from a wind forced tilt of the isopycnals at the basin scale— may be found in the ISW field (Boegman et al. 2005). This downscale energy flux has significant implications for hy- drodynamic modeling. The high-frequency waves typically have wavelengths of order 100 m (Boegman et al. 2003), much smaller than the feasible grid spacing of field-scale hydrodynamic models (Hodges et al. 2000). If the high-fre- quency waves break upon sloping topography, between 5% and 25% (Helfrich 1992; Michallet and Ivey 1999) of the incident solitary wave energy (1% to 5% of the PE o ) may be converted by diapycnal mixing to an irreversible increase in the potential energy of the water column. The generation and propagation of the progressive surge and ISW packet in natural systems with variable wind stress and complex topography are not clearly understood. Many long, narrow lakes and reservoirs are characterized by ver- tical and sloping topography at opposite ends (e.g., at dam

Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1620

Limnol. Oceanogr., 50(5), 2005, 1620–1637q 2005, by the American Society of Limnology and Oceanography, Inc.

The degeneration of internal waves in lakes with sloping topography

L. Boegman,1 G. N. Ivey, and J. ImbergerCentre for Water Research, The University of Western Australia, Crawley, Western Australia 6009, Australia

Abstract

In a laboratory study, we quantified the temporal energy flux associated with the degeneration of basin-scaleinternal waves in closed basins. The system is two-layer stratified and subjected to a single forcing event creatingavailable potential energy at time zero. A downscale energy transfer was observed from the wind-forced basin-scalemotions to the turbulent motions, where energy was lost due to high-frequency internal wave breaking along slopingtopography. Under moderate forcing conditions, steepening of nonlinear basin-scale wave components was foundto produce a high-frequency solitary wave packet that contained as much as 20% of the available potential energyintroduced by the initial condition. The characteristic lengthscale of a particular solitary wave was less than thecharacteristic slope length, leading to wave breaking along the sloping boundary. The ratio of the steepeningtimescale required for the evolution of the solitary waves to the travel time until the waves shoaled controlled theirdevelopment and degeneration within the domain. The energy loss along the slope, the mixing efficiency, and thebreaker type were modeled using appropriate forms of an internal Iribarren number, defined as the ratio of theboundary slope to the wave slope (amplitude/wavelength). This parameter allows generalization to the oceanographiccontext. Analysis of field data shows the portion of the internal wave spectrum for lakes, between motions at thebasin and buoyancy scales, to be composed of progressive waves: both weakly nonlinear waves (sinusoidal profilewith frequencies near 1024 Hz) and strongly nonlinear waves (hyperbolic–secant-squared profile with frequenciesnear 1023 Hz). The results suggest that a periodically forced system may sustain a quasi-steady flux of 20% of thepotential energy introduced by the surface wind stress to the benthic boundary layer at the depth of the pycnocline.

The turbulent benthic boundary layers located at the pe-rimeter of stratified lakes and oceans serve as vital pathwaysfor diapycnal mixing and transport (Ledwell and Hickey1995; Goudsmit et al. 1997; Wuest et al. 2000). Within thelittoral zone, this mixing drives enhanced local nutrient flux-es and bioproductivity (e.g., Sandstrom and Elliott 1984; Os-trovsky et al. 1996; MacIntyre et al. 1999). Indirect obser-vations suggest that turbulent benthic boundary layers areenergized by internal wave activity. In lakes, the turbulenceoccurs as a result of (1) bed shear induced by basin-scalecurrents (Fischer et al. 1979; Fricker and Nepf 2000; Glooret al. 2000) and (2) breaking of high-frequency internalwaves upon sloping topography at the depth of the metal-imnion (Thorpe et al. 1972; MacIntyre et al. 1999; Michalletand Ivey 1999). While the former process is relatively wellunderstood, the latter remains comparatively unexplored.

In lakes where the effects of the Earth’s rotation can beneglected, the internal wave weather may be characterized

1 To whom correspondence should be addressed. Present address:Coastal Engineering Laboratory, Department of Civil Engineering,Queen’s University, Kingston, Ontario K7L 3N6, Canada.

AcknowledgmentsThe authors thank K. G. Lamb, B. R. Sutherland, E. J. Hopfinger,

J. J. Riley, G. W. Wake, and two anonymous reviewers for com-ments and/or discussions on this work. We also thank Herve Mich-allet for supplying the laboratory data shown in Fig. 10. The LakePusiano data were collected by the Field Operations Group at CWRand the Italian Water Research Institute—in particular we thankDiego Copetti. The wavelet software was provided by C. Torrenceand G. Compo.

L.B. was supported by an International Postgraduate ResearchScholarship and a University Postgraduate Award. This researchwas funded by the Australian Research Council and forms Centrefor Water Research reference ED 1665-LB.

by the relative magnitudes of the period Ti of the horizontalmode-one internal seiche and the characteristic time scale ofthe surface wind stress, Tw (see Stevens and Imberger 1996,their table 1 for typical values). When Tw . Ti /4, an ap-proach to quasi-equilibrium is achieved, with a general ther-mocline tilt occurring over the length of the basin (Spigeland Imberger 1980; Stevens and Imberger 1996, and Fig.1b, c). From this initial condition, the resulting internal wavefield may be decomposed into a standing seiche, a progres-sive nonlinear surge, and a dispersive high-frequency inter-nal solitary wave (ISW) packet (Boegman et al. 2005). Forsome lakes, Tw K Ti, and the progressive surge is believedto originate as a wind-induced wave of depression at the leeshore (see Farmer 1978, and Fig. 1d). This wave of depres-sion will progress in the windward direction, steepening asit travels, and eventually forming a high-frequency ISWwave packet. In the absence of sloping topography, as muchas 20% of the available potential energy (PEo)—resultingfrom a wind forced tilt of the isopycnals at the basin scale—may be found in the ISW field (Boegman et al. 2005). Thisdownscale energy flux has significant implications for hy-drodynamic modeling. The high-frequency waves typicallyhave wavelengths of order 100 m (Boegman et al. 2003),much smaller than the feasible grid spacing of field-scalehydrodynamic models (Hodges et al. 2000). If the high-fre-quency waves break upon sloping topography, between 5%and 25% (Helfrich 1992; Michallet and Ivey 1999) of theincident solitary wave energy (1% to 5% of the PEo) maybe converted by diapycnal mixing to an irreversible increasein the potential energy of the water column.

The generation and propagation of the progressive surgeand ISW packet in natural systems with variable wind stressand complex topography are not clearly understood. Manylong, narrow lakes and reservoirs are characterized by ver-tical and sloping topography at opposite ends (e.g., at dam

Page 2: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1621Internal waves in closed basins

Fig. 1. (a) Schematic diagram of the experimental facility (notto scale). WG denotes ultrasonic wave gauge. See Experimentalmethods section for a detailed description. (b) Initial condition at t5 0 with upwelling at the slope. (c) Initial condition at t 5 0 withdownwelling at the slope. (d) Initial condition for the experimentsby Michallet and Ivey (1999). (b–c, and d) Representative of Tw .Ti /4 and Tw , Ti /4, respectively.

walls and river mouths, respectively). The sloping bathym-etry is believed to favor breaking of high-frequency waves(Thorpe et al. 1972; Hunkins and Fliegel 1973) as well asincreasing the production of bed shear (Mortimer and Horn1982), thus introducing significance to both the direction andmagnitude of wind-stress, which sets up the initial thermo-cline displacement. For a complete understanding of the sys-tem, this directionality must be coupled with the Tw, Ti, andthe steepening timescale, Ts (defined following).

In this study, laboratory experiments and field observa-tions are used to examine the dynamics of the large-scalewave-degeneration process in a rectangular, stratified basinwith sloping topography. Our objectives are to quantify theenergy loss from the breaking of the ISW packets as theyshoal and to cast the results in terms of parameters that areexternal to the evolving sub–basin-scale flow (e.g., windspeed and direction, boundary slope, quiescent stratification,etc.). This will facilitate engineering application and param-eterization into field-scale hydrodynamic models. We firstpresent the relevant theoretical background. The laboratoryexperiments are then described, followed by a presentationof results. Finally, the results are summarized and placedwithin the context of what is presently known about the en-ergetics of stratified lakes.

Theoretical background

During the summer months, a stratified lake will typicallypossess a layered structure consisting of an epilimnion, me-

talimnion, and hypolimnion. If the vertical density gradient isabrupt through the metalimnion, the lake may be approxi-mated as a simple two-layer system of depth h1 and densityr1 over depth h2 and density r2, where H 5 h1 1 h2 is thetotal depth and L denotes the basin length (e.g., Heaps andRamsbottom 1966; Thorpe 1971; Farmer 1978). Internalwaves may be initiated within a stratified lake by an externaldisturbance, such as a surface wind stress, t (e.g., Fischer etal. 1979, p. 161). This stress advects surface water toward thelee shore, thus displacing the internal layer interface througha maximum excursion ho as measured at the ends of the basin.The excursion is dependent on the strength and duration ofthe wind event. A steady-state tilt of the interface is achievedfor Tw . Ti/4, allowing ho to be expressed in terms of theshear velocity u* 5 Ït/ro as

2Lu*h ø (1)o g9h1

where ro is a reference density and g9 5 g(r2 2 r1)/r2 is thereduced gravity at the interface (see Spigel and Imberger1980; Monismith 1987).

The internal response of the waterbody to a forcing eventcan be gauged by the ratio of the wind-disturbance force tothe baroclinic restoring force (Spigel and Imberger 1980).Thompson and Imberger (1980) quantified this force withthe Wedderburn number, which is given for our two-layersystem in terms of Eq. (1) as

ho21W 5 (2)h1

Weak initial disturbances (W21 , 0.3) will excite a stand-ing seiche that is well described by the linear wave equation(Mortimer 1974; Fischer et al. 1979; Boegman et al. 2005).The periods of this seiche for a two-layer system are

2L(n)T 5 (3)i nco

where n 5 1,2,3, etc., denotes the horizontal mode (hereinthe fundamental timescale; Ti, without superscript, is used torepresent the gravest mode, where n 5 1) and co 5Ï(g9h1h2)/(h1 1 h2) is the linear long-wave speed.

Moderate forcing (0.3 , W21 , 1.0) results in the devel-opment of a nonlinear surge and dispersive solitary wavepacket (Thorpe 1971; Horn et al. 2001). The temporal de-velopment of the surge may be quantified by the nonlinearityparameter defined by Boegman et al. (2005) as

aa 3h zh 2 h zo 1 2; (4)c 2 h ho 1 2

where we have used the wave amplitude scaling a ; ho (e.g.,Horn et al. 1999) and the nonlinear coefficient a 5 (3/2)co(h1

2 h2)/(h1h2) (Djordjevic and Redekopp 1978; Kakutani andYamasaki 1978). If the interface is at middepth, a vanishes,steepening cannot occur, and there is no production of sol-itary waves. As the progressive nonlinear surge steepens, itslength scale decreases until nonhydrostatic effects becomesignificant and the wave is subject to dispersion (see Ham-

Page 3: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1622 Boegman et al.

mack and Segur 1978). This occurs as t → Ts, the steepeningtimescale (Horn et al. 2001)

LT 5 (5)s aho

Steepening is eventually balanced by dispersion and thesurge degenerates into a high-frequency ISW packet. Internalsolitary waves may be modeled to first order by the weaklynonlinear Korteweg–de Vries (KdV) equation

3]h ]h ]h ] h5 c 1 ah 1 b 5 0 (6)o 3]t ]x ]x ]x

where h(x,t) (positive upward) is the interfacial displacementand the dispersive coefficient b 5 (1/6)coh1h2. A particularsolution to Eq. (6) is the solitary wave equation (Benney1966),

x 2 ct2h(x 2 ct) 5 a sech (7)1 2l

where the phase velocity, c, and horizontal length scale aregiven by

1c 5 c 1 aa (8a)o 3

12b2l 5 (8b)

aa

Note the functional dependence between l and a for thenonlinear waves.

Breaking of high-frequency ISWs may occur if the wavespropagate along the density interface into coastal regionswith sloping topography. Off-shore, the interface is typicallypositioned such that h1 , h2 and a , 0, resulting in ISWsof depression. As a bounded slope is approached, the shoal-ing waves will encounter a turning point, where h2 5 h1 anda 5 0. Beyond the turning point, h1 . h2 and a . 0, causingthe ISW to change polarity and become waves of elevation.Analytical KdV theories (see Miles 1981 for a review) havebeen extended to model the evolution and propagation ofsolitary waves when the topography and background floware slowly varying (e.g., Lee and Beardsley 1974; Djordjevicand Redekopp 1978; Zhou and Grimshaw 1989). Horn et al.(2000) further extended these theories to a strong space–timevarying background, but propagation through the turningpoint and wave breaking were necessarily avoided. First-order analytical models do not allow transmission of theISWs through the singular point at a 5 0.

The location of the turning point may be easily obtainedfor a quiescent two-layer flow with sloping topography inthe lower layer. Define the thickness of the lower layer alongthe slope in x as

Hh (x) 5 H 2 h 2 x (9)2 1 1 2Ls

where Ls denotes the slope length and x 5 0 at the toe ofthe slope. The location of the turning point on the slope isgiven in nondimensional form by letting h2(x) 5 h1

x h 2 h2 15 (10)L Hs

As a wave travels along the slope, it steepens, a increases,and the streamlines approach vertical. During steepening, themaximum horizontal fluid velocity in the direction of wavepropagation umax will increase more rapidly than c and alimiting amplitude may be achieved where the velocities inthe wave crest become equal to the phase velocity. Here, thislimit is defined as the breaking limit and the location on theslope where this limit is observed is the breaking point.

The type of internal wave breaking that occurs at thebreaking point may be inferred by analogy to surface break-ers. Of the continuum of surface breaker types that exists,four common classifications are used (see Komar 1976).Spilling breakers occur on mildly sloping beaches wheresteep waves gradually peak and cascade down as white wa-ter. Plunging breakers are associated with steeper beachesand waves of intermediate steepness, the shoreward face ofthe wave becomes vertical, curling over and plunging for-ward as an intact mass of water. Surging breakers occur onsteep beaches where mildly sloping waves peak up as if toplunge, but the base of the wave surges up the beach face.Collapsing breakers, which mark the transition from plung-ing to surging, begin to curl over and then collapse uponthemselves with some of the water mass surging forward upthe slope.

For surface waves, Galvin (1968) and Battjes (1974)found that the ratio of the beach slope S to the wave slope(a/l) was suitable for classifying the breaker type. This ratiois expressed as either off-shore or near-shore forms of theIribarren number j,

Sj 5 (11a)b 1/2(a /l )b `

Sj 5 (11b)` 1/2(a /l )` `

where the subscripts ` and b refer to off-shore and near-shore wave properties, respectively (Fig. 2a,b). The breakerheight ab is measured when the wave face first becomes ver-tical in the surf zone.

The dynamics of internal wave breaking on sloping to-pography have been classified according to the ratio of l`

to the slope length, Ls (e.g., Michallet and Ivey 1999; Bour-gault and Kelley 2003). This parameter retains no knowledgeof the actual boundary slope and consequently may not beused to generalize results. This is shown in Fig. 2c,d, where,in both diagrams, l`/Ls is equivalent, yet H1/Ls ± H2/Ls anddifferent breaking dynamics are expected as the internalwaves shoal. The utility of the ratio of the beach slope tothe internal wave slope has been suggested (e.g., Legg andAdcroft 2003), but a formal classification based on internalwave data has yet to be performed. Perhaps this stems fromthe difficulty in measuring the parameters ab, l`, and a` forinternal waves. It is preferable to recast j in terms of readily

Page 4: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1623Internal waves in closed basins

Fig. 2. Schematic diagram of wave and slope properties: Off-shore wave amplitude, a`, off-shore wavelength, l`, near-shorebreaker height, ab, boundary slope, S, slope-length, Ls, and slope-height, H1 and H2.

Table 1. Summary of experimental runs. The experimental variables together with the resolution with which they were determined: theslope, S, the interface depth, h1 (60.2 cm), the maximum interface displacement, h0 (60.2 cm), the horizontal mode one basin-scale internalwave period (Ti), the steepening time scale (Ts), and the density difference between the upper and lower layers, Dr ø 20 kg m23 (62 kgm23). The h0 values were measured along the vertical end wall, where the 1 and 2 symbols denote an initial displacement as measuredabove and below the quiescent interface depth, respectively (i.e., 1 as in Fig. 1c and 2 as in Fig. 1b).

Run S h1/H W21 Ti (s) Ts (s) Run S h1/H W21 Ti (s) Ts (s)

12345

3/203/203/203/203/20

0.190.180.210.290.29

20.3920.8620.9320.3820.58

130133126113113

1466661

167109

1920212223

1/101/101/101/101/10

0.190.200.190.300.30

20.2620.8120.8520.2320.50

130128130112112

2187067

283130

6789

10

3/203/203/203/203/20

0.290.470.470.480.20

20.9020.3520.7320.9010.43

113102102102128

71862413493132

242526—27

1/101/101/10

—1/10

0.270.510.51

—0.20

20.9220.2920.59

—10.15

115102102—128

662,8801,416

—379

1112131415

3/203/203/203/203/20

0.190.200.300.300.30

10.6210.8610.3410.6010.86

130128112112112

9266

19110876

2829303132

1/101/101/101/101/10

0.200.200.270.300.30

10.8511.010.2510.5010.91

128128115112112

6757

24413072

161718

3/203/203/20

0.490.500.50

10.2910.4110.82

102102102

2,997——

3334—

1/101/10

0.510.51

10.2910.66

102102—

2,8801,265

measured variables (e.g., wind speed, quiescent fluid prop-erties). This is accomplished for ISWs with a sech2 profileby noting that the wave slope a`/l` ; aho/co, resulting in

Sj 5 (12)sech 1/2(ah /c )o o

where ho is easily estimated using (1). Similarly, for sinu-soidal waves a`/l` ; fho/co giving

Sj 5 (13)sin 1/2( fh /c )o o

In the following sections, progressive sinusoidal waves areaddressed and the wave frequency, f, is discussed.

Experimental methods

The experiments were conducted in a sealed rectangularacrylic tank (600 cm long, 29 cm deep, and 30 cm wide)into which a uniform slope of either 1/10 or 3/20 was in-serted, extending the entire height of the tank and positionedat one end (Fig. 1). The tank was filled with a two-layerstratification from reservoirs of fresh and saline filtered wa-ter (0.45-mm ceramic filter). For visualization purposes, thelower layer was seeded with dye (Aeroplane Blue Colour133, 123). Prior to commencing an experiment, the tank wasrotated to the required interfacial displacement angle. Fromthis condition, the setup and subsequent relaxation from awind-stress event was simulated through a rapid rotation ofthe tank to the horizontal position, leaving the interface in-clined at the original angle of tilt of the tank. Depending onthe initial direction of rotation prior to commencing an ex-periment, the resulting inclined interface at t 5 0 could becharacterized as either upwelling on the slope (Fig. 1b) ordownwelling on the slope (Fig. 1c). The ensuing verticaldisplacements of the density interface h(x,t) were measuredusing three ultrasonic wave gauges (Michallet and Barthe-lemy 1997) distributed longitudinally along the tank at lo-cations A, B, and C (Fig. 1). The wave gauges logged datato a personal computer at 10 Hz via a 16-bit analog-to-digitalconverter (National Instruments PCI-MIO-16XE-50). Theexperimental variables considered in this study, togetherwith the resolution with which they were determined, aregiven in Table 1.

To visualize the interaction of the internal wave field withthe sloping topography, digital images with a resolution of1 pixel per mm were acquired at a rate of 5 Hz from aprogressive scan CCD camera (PULNiX TM-1040)equipped with a manual zoom lens (Navitar ST16160). In-

Page 5: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1624 Boegman et al.

Fig. 3. Video frames showing the wave field evolving from theinitial condition in (a). The surge and ISW packet are propagatingto the right in (b) and to the left in (c), (d), (f), and (g). Wavebreaking is shown to occur upon the slope. For this experiment, h1/H 5 0.29 and ho/h1 5 0.90. The apparent dye-free layer near thetank bottom is a spurious artifact of light reflection.

Fig. 4. (a) Time series and (b) continuous wavelet transformsshowing the temporal evolution of the internal surge and solitarywave packet for the case of no slope (case 1). (c,d) Same as (a) and(b) except for the initial condition of upwelling on the 3/20 slope(case 2). (e,f) Same as (a) and (b) except for the initial conditionof downwelling on the 3/20 slope (case 3). For all experiments, ho/h1 ø 0.7, h1/H ø 0.2, and Ts ø 70 s. The arrows in (c) and (e)denote the direction of wave propagation relative to the tank sche-matics shown in Fig. 1c,d, respectively.

dividual frames were captured in a LabVIEW environmentfrom a digital frame grabber (National Instruments PCI-1422) and written in real-time to disk. The tank was illu-minated with backlight from twenty-eight 12-V, 50-W hal-ogen lamps (General Electric Precise MR16; Fig. 1). Toremove flicker associated with the AC cycle, a constant DCsupply was maintained to the lamps via four 12-V car bat-teries. The images were corrected for vignetting, variationsin illumination intensity, dust, and aberrations within the op-tical system and pixel gain and offset following the methodsof Ferrier et al. (1993). This procedure required flat-fieldcomposite images of (1) the room darkened and the lens capon the camera (dark image), as well as (2) the tank filledwith dyed saline water from the reservoir (blue image), and(3) filtered tap water (white image). From the three flat-fieldimage composites, a linear regression was performed on thegray scale response at each pixel location versus the meanpixel response across the CCD array to yield slope and in-tercept arrays. The experimental images were subsequentlycorrected at each pixel by subtracting the dark image, sub-tracting the intercept array, and dividing by the slope array(e.g., Cowen et al. 2001). Note that a correction for lightattenuation was not required as a result of the backlight ge-ometry. The gray scale pixel response within each imagewas then calibrated to the fluid density by associating mea-sured densities in the fluid layers with the mean correctedpixel response of the blue and white images and mappingthe intermediate values according to the log-linear dye ver-sus concentration relationship, as determined from an incre-mental calibration procedure in a test cell.

Results

Flow field—To illustrate the qualitative features of theflow, let us concentrate on the behavior of a particular ex-periment (run 6, Table 1). From the initial condition, asshown in Fig. 3a, the flow was accelerated from rest by thebaroclinic pressure gradient generated by the initial tilted

density interface. Characteristic of a standing horizontalmode-one (H1) seiche, the lower layer moved toward thedownwelled end of the tank with a corresponding return flowin the upper layer. Progressing initially from left to right, aninternal surge and ISW packet were also evident (Fig. 3b).These reflected from the vertical end wall and progressedleft toward the slope (Fig. 3c). The ISWs subsequentlyshoaled upon the topographic slope and wave breaking wasobserved (Fig. 3d,e). Not all wave energy was dissipated atthe slope, as a longwave was reflected (Fig. 3e, right side oftank). This was reflected once again from the vertical walland a secondary breaking event was observed (Fig. 3g). Af-ter the secondary breaking event, the internal wave field wasrelatively quiescent (Fig. 3h).

The frequency content of the internal surge and ISW pack-et, as well as the effects of the directionality associated withthe initial forcing condition (i.e., upwelling or downwellingon the slope), were examined through a time–frequencyanalysis carried out using continuous wavelet transforms(Torrence and Compo 1998). Results from wavegauge B, anodal location for the H1 seiche, are shown for three rep-resentative experiments (Fig. 4), where the forcing ampli-

Page 6: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1625Internal waves in closed basins

Fig. 5. Schematic illustration showing the evolution of the spa-tial wave profile as observed in the laboratory. (a–f) Evolution fromthe initial condition in Fig. 1c, g–l. Evolution from the initial con-dition in Fig. 1b. Illustrations are representative of Ti /2 time inter-vals. Arrows denote wave-propagation direction.

tude, ho/h1 ø 0.7, and layer thickness, h1/H ø 0.2, weresimilar. Case 1 is a reference case with no slope (Fig. 4a,b).The majority of the energy was initially found in the pro-gressive surge with a period of Ti /2. For t/Ts $ 1, the energyappeared to be transferred directly from the surge to solitarywaves with a period near Ti /16; there was no evidence ofenergy cascading through intermediate scales (cf. Horn et al.2001; Boegman et al. 2005). The ISW packet persisted untilt/Ts . 6, the wave period gradually increasing from Ti /16 toTi /8 as the rank-ordered waves in the packet separate ac-cording to Eq. (8a). The waves were eventually dissipatedby viscosity. In case 2, the initial condition of upwelling onthe slope (Fig. 4c,d) is as shown in Fig. 1b. The ISW packetinitially evolved in the same manner as case 1, for a smalltime. However, over 2 , t/Ts , 3, the fully developed ISWpacket interacted with the sloping topography and rapidlylost its high-frequency energy to wave breaking (see Fig. 3).The reflected long wave, shown progressing to the right att/Ts ø 3, traveled the tank length and continued steepening.A second ISW packet was observed to develop and wasobserved to break over 4 , t/Ts , 5. Case 3 is the initialcondition of downwelling on the slope (Fig. 4e,f), as shownin Fig. 1c. In this experiment, the ISW packet initially in-teracted with the slope at t/Ts ø 1. At this time, the wavepacket was in the early stages of formation and only a weakbreaking event, with little apparent energy loss from thewave, was observed. The reflected wave packet then contin-ued to steepen as it propagated to the right and reflected offthe end wall. Again, a second breaking event was observedover 3 , t/Ts , 4, with more significant energy loss fromthe wave. From these results in closed systems, it is clearthat the nature of the surge and ISW packet life history isdependent not only on the bathymetric slope but also on thedirection of the forcing relative to the position of the topo-graphic feature.

This concept is illustrated in Fig. 5, where visualizationof the laboratory experiments (e.g., Fig. 3) has allowed thewave degeneration process, from the initial conditionsshown in Fig. 1b,c, to be conceptually depicted. A longwaveinitially propagates from the upwelled fluid volume (Fig. 5band h). This wave steepens as it travels, thus producing high-frequency ISWs. The wave groups reflect from the verticalwall with minimal loss and break as they shoal along thebathymetric slope. The breaking process is expected to begoverned by j.

Internal solitary wave energetics—In the preceding sec-tion, a qualitative description of the wave–slope dynamicsin a closed basin was presented. To quantify the temporalevolution of the ISW energy and the energy loss from thesewaves as they interact with the topographic boundary slope,we applied the signal processing methods of Boegman et al.(2005). This technique requires the energy in the variousinternal modes to be discrete in frequency space (as dem-onstrated in Fig. 4), thus allowing the individual signals tobe isolated through selective filtering of the interface dis-placement time series. By assuming an equipartition betweenkinetic and potential forms of wave energy, the energy inthe surge (ENS) and ISW wave packets (EISW) was deter-mined from the filtered time series as the integral of the

signal over the wave/packet period. Note that net downscaleenergy flux through wave breaking at the boundary may notbe quantified by this method because we cannot account forthe observed upscale energy flux to reflected longwaves. Re-sults are presented in Fig. 6, showing contours of ENS/PEo

(Fig. 6a) and EISW/PEo (Fig. 6b–f). In each panel, the verticalaxis (aho/co) indicates the relative magnitudes of the linearand nonlinear components of the internal wave field (fromEq. (4)), while the horizontal axis (Ti) reveals how the sys-tem evolves in time. The times at which the solitary wavepacket shoals upon the sloping beach and Ts are also indi-cated (see caption).

For case 1, where there is no slope, the surge energy (ENS/PEo) is shown to increase during an initial nonlinear steep-ening phase (0 , t # Ts) and retain up to 20% of the PEo

at t ø Ts (Fig. 6a). When t . Ts, the ENS/PEo decreased fromthis maximum, with a corresponding increase in ISW energy,until eventually all of the surge energy was transferred tothe solitary waves with little loss (Fig. 6b). In the absence

Page 7: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1626 Boegman et al.

Fig. 6. General evolution of internal wave energy normalizedby the PEo introduced at the initial condition: (a) Surge energy (i.e.,ENS/PEo) and (b) ISW energy (i.e., EISW/PEo) for the case of noslope. The contours in (a) and (b) are least squares fit to the data(see Boegman et al. 2004, 2005). ISW energy for the case of (c)upwelling along the 3/20 slope and (d) upwelling along the 1/10slope. ISW energy for the initial condition of (e) downwelling alongthe 3/20 and (f) downwelling along the 1/10 slope. Contours arepresented as a percentage of the PEo introduced at t 5 0 and arecompiled using the data in Table 1 and Boegman et al. (2005). Thecontour interval is 5%. The ratio Ts /Ti is indicated (solid line) fora particular aho/co. The times at which the solitary wave packetshoals upon the sloping beach are also shown (dashed line).

Fig. 7. Images of mixing as the layer interface advances alongthe slope for t . 0. Data from experimental run 6. The horizontalwindow length is 1 m and the aspect ratio is 1:1. The upper layerdensity r1 ø 1,000 kg m23, and the lower layer density r2 ø 1,020kg m23.

of sloping topography, the energy in the ISW packet wasultimately lost to viscosity on time scales of order 3Ti to 5Ti.

Figure 6c,d shows results for case 2, upwelling on bound-ary slopes of 3/20 and 1/10, respectively. In both panels, theevolution of EISW/PEo was initially analogous to case 1 withno slope; however, now the waves shoaled after traveltimesof 0.5Ti, 1.5Ti, and 2.5Ti. The first wave/slope interactionoccurred at 0.5Ti, prior to the development of any high-fre-quency ISWs (i.e., t , Ts). At this time, there was no dis-cernible breaking of the basin-scale surge (with a mild waveslope) on the relatively steep boundary slope. The characterof the second wave-slope interaction was dependent on themagnitude of aho/co. For aho/co . 0.5, we find Ts , 1.5Ti,which results in nearly all ISW energy (as much as 25% ofPEo) being lost near the boundary or reflected during the1.5Ti shoaling event. For some experiments where aho/co ø0.5, secondary breaking events were observed at 2.5Ti. Ex-perimental visualization revealed that these events transpiredwhen the ISW packet had not fully developed before theprimary breaking event (i.e., Ts ø 1.5Ti); hence, subsequentpacket development occurred prior to 2.5Ti.

In Fig. 6e,f, results are shown for case 3, the initial con-dition of downwelling at boundary slopes of 3/20 and 1/10,respectively. Again, the evolution of EISW/PEo is initiallysimilar to the case with no slope but the waves then shoaledaround Ti and 2Ti. By t ø Ti, wave steepening had begun,

EISW/PEo energy levels were about 5–10%, and this energywas lost from the ISW field at the boundary. As with thecase of upwelling at the slope, for aho/co ø 0.5, a secondarybreaking event may occur at 2Ti.

It is clear from Fig. 6 that the forcing direction (upwellingor downwelling) determines the travel time of the ISW pack-et until it shoals, as a multiple of Ti /2. Similarly, the growthof the ISW packet is governed by Ts. The key nondimen-sional parameter, relative to the traveltime, is Ts/Ti (Fig. 6).From Eqs. (3) and (5), this parameter is

T 1 h hs 1 25 (14)T 3 (h 2 h )hi 1 2 o

We note that in Fig. 6c,d, the maximum EISW/PEo ø 20%occurred at aho/co ø 0.5; whereas in Fig. 6b, the maximumEISW/PEo ø 20% occurred at aho/co ø 1. Experimental vi-sualization suggests that the slight reduction in wave-fieldenergy as aho/co → 1 for the experiments with upwellingalong the topographic slope may have resulted from en-hanced energy transfer to mixing and dissipation as the den-sity interface accelerated along the slope from the initial restcondition (Fig. 7). This mixing did not occur during the ex-periments with vertical end walls.

Page 8: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1627Internal waves in closed basins

Breaker observations—Following the success of classi-fying surface breakers according to the Iribarren number, wefound it instructive to test these classifications for internalwave breaking. Measurements of a` and l` were obtainedfrom the wave-gauge data at wave-gauge B. Images fromthe CCD camera were used to estimate ab and classify thebreaker type. In Figs. 8 and 9, internal waves are shown tospill, plunge, and collapse, suggesting that a form of theIribarren classification may be suitable for internal wavebreaking.

Spilling breakers were observed on the milder slope (S 51/10), when the wave steepness is high (aho/co . 0.8). Twoexamples of spilling breakers are presented in Fig. 8a–f andg–l. In both examples, the rear face of each ISW in thepacket breaks as a wave of elevation in a spilling manner.Mixing and dissipation appear to be very localized and onoccasion (e.g., Fig. 8c), baroclinic motions shear off the crestof the wave. These observations are similar to those by Hel-frich (1992, his figs. 3–5).

Plunging breakers were observed on both slopes for mod-erately steep waves (0.4 , aho/co , 0.8). Plunging occurredprimarily during the breaking events at 1.5Ti in Fig. 6c,d.The initial condition of upwelling on the slope allowed along travel time (1.5Ti) and a fully developed wave packet.Recall that, in these breaking events, up to 20% of the PEo

is lost from the high-frequency ISW field. Two examples ofspilling breakers (Fig. 8m–r and s–x) suggest these eventsare more energetic than the spilling class. In each example,an intact mass of fluid from the rear face of each solitarywave of depression was observed to plunge foreward (e.g.,Fig. 8o and u), becoming gravitationally unstable and form-ing a core of mixed fluid, which propagates upslope. Theseobservations are similar to numerical results by Vlasenkoand Hutter (2002, their fig. 6).

Collapsing breakers were also observed to occur on bothboundary slopes, when the incident wave slope was milderthan that required for plunging. Two examples of collapsingbreakers (Fig. 9a–f and g–l) show that the breaking eventswere less energetic than plunging and were characterized bya mass of fluid beginning to plunge forward (e.g., Fig. 9dand i), and then collapsing down upon itself (e.g., Fig. 9e,and j–k). Compared with plunging breakers, little mixingoccurred. These observations are similar to those by Mich-allet and Ivey (1999, their fig. 3).

For experiments where h1 5 h2, a 5 0, and first-orderKdV type solitary waves are prohibited from forming. Suf-ficiently strong forcing (ho/h1 . 0.4) may, however, generatesupercritical flow conditions and an internal undular bore(see Horn et al. 2001). The large baroclinic shear associatedwith this forcing was capable of generating Kelvin–Helm-holtz instability, both in the tank interior (Horn et al. 2001,their fig. 8) and at the boundary (Fig. 9m–r and s–x). Break-ing of the undular waves through shear instability was ob-served to result in strong mixing of the water column (Fig.9r,x). These observations are qualitatively similar to oceanicfield observations by Moum et al. (2003, their fig. 14) andOrr and Mignerey (2003, their fig. 9).

Breaker classification and reflection coefficient—Thebreaker visualizations, as presented above, allow classifica-

tion of the breaker type in terms of j and the measuredreflection coefficient, R 5 Er/Ei, where Er and Ei are theenergy in the reflected and incident wave packets, respec-tively, calculated from the wave-gauge B signal. The signalis not filtered in order to allow for the observed upscaleenergy transfer from the breaking high-frequency wavepacket to a reflected longwave, which scales as the packetlength (reflected long-wave energy would be removed by thefilter employed in Fig. 6). Our experimental methodologydid not permit calculation of the mixing efficiency G for aparticular wave breaking event. However, G may be obtainedfrom published observations of breaking lone solitary waves(Michallet and Ivey 1999, their fig. 10), when expressed asa function of j.

In Fig. 10a–c, we plot R versus jb, j`, and jsech, respec-tively. For small j, the wave slope is steep relative to S andthe wave propagation over the slope approaches that of awave in a fluid of constant depth, spilling breakers are ob-served, R → 0, and viscosity dominates, with minimal waveenergy converted to an increase in the potential energy ofthe system through diapycnal mixing (Fig. 10d,e). Converse-ly, for waves with very large j, the time scale of wave–slopeinteraction is small, R → 1, and the collapsing breakers againinduce minimal mixing. The inertia of these waves appearsunable to overcome the stabilizing effect due to buoyancy.Intermediate to these extremes, plunging breakers developwith gravitational instabilities that drive mixing efficiencies,peaking near 25%. As the incident waves contain as muchas 25% of the PEo, it implies that 6% of the PEo may beconverted by diapycnal mixing to an irreversible increase inpotential energy. These results are consistent with those forlone solitary waves by Michallet and Ivey (1999), whichhave been recast in terms of j and are also presented in Fig.10b,c. The present results, however, have distinctly lower Rvalues. This may be attributed to the observation that de-caying turbulence and wave reflection associated with theleading wave generally interferes with subsequent waves inthe packet, increasing fluid straining and therefore dissipa-tion within the breaker. For this same reason, the breakertype classifications strictly apply to only the leading wavein each packet. Some differences, between these results andthose of Michallet and Ivey (1999), might also arise fromthe presence of a no-slip boundary condition at the top sur-face (Redekopp pers. comm.).

Interpretation of our results, in terms of the mixing modelby Ivey and Imberger (1991), suggests that, in the spillingand collapsing regions, mixing is suppressed by viscosityand buoyancy, respectively; consequently G , 15%. In theplunging region, inertia dominates, suggesting that the scaleof the most energetic overturns approaches the Ozmidovscale. As a result, the potential energy available for mixingis maximized and G . 15%. Such large-scale overturning isalso observed for Kelvin–Helmholtz billowing (Imbergerand Ivey 1991) and critical internal wave breaking (Ivey andNokes 1989).

The breaking point—Now that breaker classificationguidelines have been established and the energy loss at theslope has been quantified, it is useful to estimate the positionalong the slope where breaking occurs (breaking point).

Page 9: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1628 Boegman et al.

Fig. 8. False color images of spilling breakers: (a–f) run 28, (g–l) run 32. False color images of plunging breakers: (m–r) run 6, (s–x)run 12. The horizontal window length is 1 m and the aspect ratio is 1:1. The upper layer density r1 ø 1,000 kg m23, and the lower layerdensity r2 ø 1,020 kg m23.

Page 10: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1629Internal waves in closed basins

Fig. 9. False color images of collapsing breakers: (a–f) run 14, (g–l) run 11. False color images of Kelvin–Helmholtz breakers: (m–r)run 17, (s–x) run 18. The horizontal window length is 1 m and the aspect ratio is 1:1. The upper layer density r1 ø 1,000 kg m23, andthe lower layer density r2 ø 1,020 kg m23.

Page 11: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1630 Boegman et al.

Fig. 10. Reflection coefficient (R), mixing efficiency (G), and breaker type classified accordingto the various forms of the Iribarren number: (a) Breaker type and R versus jb, (b) breaker typeand R versus j`, (c) breaker type and R versus jsech, (d) breaker type and G versus j`, and (e) breakertype and G versus jsech. The dashed lines demarcate the breaker classifications and are inferred in(d) and (e) from (b) and (c), respectively. The G observations in (d) and (e) are from Michallet andIvey (1999).

Most studies to date have been concerned with the mechan-ics (e.g., Kao et al. 1985; Lamb 2002; Legg and Adcroft2003) and location (e.g., Helfrich and Melville 1986; Vla-senko and Hutter 2002) of wave breaking upon the slope–shelf topography characteristic to the coastal ocean. Com-paratively fewer studies have considered the propagation andbreaking of ISWs over a topographic ridge (e.g., Farmer and

Armi 1999; Sveen et al. 2002). Relevant to the present study,shoaling ISWs on closed slopes have been examined in thelaboratory by Wallace and Wilkinson (1988) and Helfrich(1992). However, in the former study, a turning point wasnot encountered as h2 , h1. Helfrich (1992) determined abreaking criterion in terms of the undisturbed lower-layerdepth at the breaking point hb and the distance from the

Page 12: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1631Internal waves in closed basins

Fig. 11. The breaking point (criterion) as a function of the un-disturbed lower layer depth (hb). Li is the distance from the begin-ning of the slope to the undisturbed interface–slope intersection.The experiments with slope, S 5 0.034, 0.050, and 0.067 are fromHelfrich (1992, his fig. 6).

Fig. 12. Theoretical turning point versus measured breakingpoint. Various breaker types are shown: spilling, collapsing, andplunging. Error bars denote uncertainty in determining the breakingpoint due to parallax (shows maximum and minimum position).

Fig. 13. Map showing the position of Lake Pusiano in Italy(48.808N, 9.278E) and the lake bathymetry. A thermistor chain (in-dividual sensors every 0.75 m between depths of 0.3 and 24.3 m)and weather station (2.4 m above the surface) were deployed atstation T during July 2003. Data were sampled at 10-s intervals.Isobaths are given in meters.

beginning of the slope to the undisturbed interface–slope in-tersection, Li. In Fig. 11, the breaking points for the presentexperiments (as observed from the CCD camera) are shownto complement these results. The tendency for a`/hb to in-crease as l`/Li decreases is found to also be valid for steeperslopes, supporting the observation that larger waves tend topropagate into slightly shallower water (relative to their ini-tial amplitude) before breaking (Helfrich 1992). An increasein bed-slope with l`/Li is also evident. The excursion of theinterface along the slope and enhanced interfacial shear as-sociated with the basin-scale seiche are likely responsiblefor the observed scatter in the breaking points of the presentstudy, relative to (Helfrich 1992). The line of best-fit to thedata

a 0.14` 5 2 0.03 (15)0.52h (l /L )b ` i

allows generalization of the results.An analytical breaking criterion was derived from Eq. (6),

using the variable coefficients a and b that are dependenton the local fluid depth (e.g., Djordjevic and Redekopp1978). This first-order weakly nonlinear form of KdV theorydoes not permit wave breaking (Whitham 1974) and is alsolimited to relatively steep slopes, where breaking occurs rap-idly; thus, higher order nonlinear effects predominate over arelatively short transition zone. For details of the analysis,the reader is referred to Boegman (2004).

The observations presented thus far show the internal sol-itary waves of depression to ultimately break as waves ofelevation. Before breaking, these waves must change polar-ity by passing through the turning point (beyond which a .0). This is shown to occur in Fig. 12, where the observedlocation of the breaking point is consistently after the the-oretical prediction for the location of the turning point. Notethat Eq. (10) is strictly appropriate for an undisturbed two-layer stratification and does not account for the excursion ofthe interface along the slope and enhanced interfacial shearassociated with the basin-scale seiche.

Field observations

Lake Pusiano—To assess the applicability of our results,we applied our analysis to some recent observations from alake with suitable field data and topography (Lake Pusiano,Italy; Fig. 13). This small, seasonally stratified lake, withnear vertical and gradually sloping bathymetry on the northand south shores, respectively, is subject to pulses of highwind stress from the northerly direction (Fig. 14a). The ep-isodic wind events, on days 203 and 205, introduce energyto the lake at the basin scale. For these events, Tw . Ti /4and a general thermocline tilt is expected. Using the ob-served wind velocity, W21 5 ho/h1 was calculated (Table 2)from Eqs. (1) and (2). Following Stevens and Imberger(1996), the wind stress during each wind event was averagedand integrated over Ti /4. The response of the internal wavefield to these wind events, as observed at station T, is con-

Page 13: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1632 Boegman et al.

Fig. 14. Observations of wind speed and direction and isotherm displacement from Lake Pus-iano at T. (a) Mean hourly wind direction (dots), and 10-min average wind speed (solid line),corrected to 10 m above the surface; (b) isotherms at 28C intervals calculated through linear inter-polation of thermistor data at T (10-min average); (c) detail of shaded region in (b) (1-min average).Note change of time base. The bottom isotherm in (b) and (c) are 88C and 108C, respectively. (d)Normalized instantaneous energy calculated as cog(r2 2 r1)h from the displacement of the 208C2

20

isotherm (h20) in (c) and the integral of this quantity shown as a fraction of the PEo over Ti /2intervals as shown (see Table 2). PEo calculated assuming a linear h20 tilt of maximum excursiongiven by Eq. (1). (e) Same as (d) except for the layer interface from experimental run 31 ratherthan h20. Letters b through f denote corresponding panels in Fig. 5.

sistent with other published field observations (e.g., Hunkinsand Fliegel 1973; Farmer 1978; Boegman et al. 2003). Spe-cifically, the generation and subsequent rapid dissipation ofhigh-frequency internal waves was observed (Fig. 14b). Todirectly compare these observations to the laboratory results,an approximation to a two-layer stratification was made byseparating the epilimnion from the hypolimnion at the 208Cisotherm (h20). The nonlinear waves were isolated from con-tamination, due to motions at the basin scale and buoyancy

frequency, by low-pass filtering and detrending h20. The totalenergy was then integrated over the period of a progressivewave packet Ti /2

t 1T /2o i

2E 5 c g(r 2 r ) h (t) dt (16)total o 2 1 E 20

to

(see Boegman et al. 2005). The observed temporal evolutionof h and Etotal (Fig. 14c,d) are similar to laboratory data2

20

Page 14: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1633Internal waves in closed basins

Tabl

e2.

Tabu

late

dli

stof

fiel

dob

serv

atio

nsto

geth

erw

ith

the

calc

ulat

edpa

ram

eter

sus

edin

this

stud

y.W

ith

the

exce

ptio

nof

Lak

eP

usia

no,

the

obse

rvat

ions

ofh

0ar

efr

om(H

orn

etal

.20

01,

thei

rta

ble

1).

Par

amet

ers

calc

ulat

eddi

rect

lyfr

omob

serv

atio

nsin

Lak

eP

usia

no(F

ig.

14)

are

deno

ted

by‡.

Ran

dG

wer

ees

tim

ated

from

Fig

.10

b–e.

N/A

deno

tes

apa

ram

eter

that

isno

tap

plic

able

.T

he—

sym

bol

deno

tes

that

data

,w

hile

appl

icab

le,

wer

eno

tav

aila

ble.

Loc

atio

nS

h 1 (m)

h 2 (m)

a ` (m)

f(H

z)l

`

(m)

h0

(m)

j sec

h†or

j sin*

j `R (%

)G (%

)S

ourc

e

Bod

ense

eK

oote

nay

Lak

eL

ake

Pus

iano

(day

203)

Lak

eP

usia

no(d

ay20

5)K

oote

nay

Lak

e

1/10

01/

201/

501/

501/

20

30 30 7 7 30

75 65 17 17 65

10 5 2 2 5

2.3 3

102

5

1.43

102

4

1.53

102

4

2.73

102

4

2.83

102

4

6,50

02,

520

2,13

01,

150

1,26

0

— 15 3.4‡

3.0‡

15

—*

0.6*

0.5*

0.4*

0.5*

0.3

1 0.6

0.4

0.8

20 50 28‡

26‡

40

5–10

5–25

15–2

05–

1520

–25

Unp

ubl.

data

(200

1)W

iega

ndan

dC

arm

ack

(198

6)F

ig.

14F

ig.

14W

iega

ndan

dC

arm

ack

(198

6)S

ulu

Sea

Loc

hN

ess

Bab

ine

Lak

eB

abin

eL

ake

Bod

ense

e

1/11

1/20

1/30

01/

300

1/10

0

170 35 20 20 25

3,30

011

514

714

7 75

65 12 10 15 20

3.7 3

102

4

4.23

102

4

4.23

102

4

4.23

102

4

5.83

102

4

6,30

088

046

046

025

0

N/A

12 10 10 —

—†

0.08

†0.

004†

0.00

4†—

0.9

0.4

0.02

0.02

0.04

40 20 — — —

15–2

05–

15— — —

Ape

let

al.

(198

5)T

horp

eet

al.

(197

2)F

arm

er(1

978)

Far

mer

(197

8)U

npub

l.da

ta(2

001)

Nor

thW

est

She

lfN

orth

Wes

tS

helf

Str

ait

ofG

ibra

ltar

Lak

eB

iwa

Ore

gon

She

lf

3/2,

000

3/2,

000

N/A

1/20 —

43 48 40 12 7

80 75 960 38 140

30 35 40 12 21

8.3 3

102

4

8.33

102

4

1.13

102

3

1.53

102

3

1.63

102

3

910

830

— 460

N/A

N/A

N/A

12 —

—†

—†

—†

0.05

†—

0.01

0.01 — 0.3 —

— — —10

–20

— — — 5–10

Hol

low

ay(1

987)

Hol

low

ay(1

987)

Zie

genb

ein

(196

9)B

oegm

anet

al.

(200

3)S

tant

onan

dO

stro

vsky

(199

8)S

enec

aL

ake

Mas

sach

uset

tsB

ayM

assa

chus

etts

Bay

Sco

atia

nS

helf

1/16

1/20

01/

200

1/20

0

20 7 7 50

68 75 75 100

20 10 10 30

1.7 3

102

3

2.13

102

3

2.83

102

3

5.33

102

3

210

422

317

1,10

0

10 N/A

N/A

N/A

0.09

†—

†—

†—

0.2

0.03

0.03

0.01

10–2

0— — —

5–10

— — —

Hun

kins

and

Fli

egel

(197

3)H

alpe

rn(1

971)

Hal

pern

(197

1)S

ands

trom

and

Ell

iott

(198

4)S

t.L

awre

nce

Lak

eK

inne

ret

Lak

eK

inne

ret

1/25 N/A

N/A

9 15 12

61 17 20

16 2 3

6.5 3

102

3

6.33

102

3

6.33

102

3

124 30 30

N/A

N/A

N/A

0.03

†N

/AN

/A

0.1

N/A

N/A

6–8

— —

— — —

Bou

rgau

ltan

dK

elle

y(2

003)

Boe

gman

etal

.(2

003)

Boe

gman

etal

.(2

003)

Page 15: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1634 Boegman et al.

from an experiment (run 31) with comparable initial andboundary conditions as shown in Fig. 14e (downwelling onslope, h1/H ø 0.3 and 0.4 , ho/h1 , 0.5). During 0 , t ,0.5Ti, the field and laboratory data show Etotal /PEo ø 20%and Etotal /PEo ø 40%, respectively. The lower energy levelsin the field appear to be attributed to the absence of a pro-gressive nonlinear surge, which, in the laboratory, carriedthe high-frequency waves. During 0.5Ti , t , 1.5Ti, thehigh-frequency oscillations in both systems contain approx-imately 15% of Etotal /PEo, and this energy rapidly decreasesto approximately 5% as t . 1.5Ti.

Figure 5 allows the wave degeneration process in LakePusiano to be conceptually depicted. The initial condition, anortherly wind and sloping southern shore, was as shown inFig. 5a. Upon termination of the wind stress, the systemevolves schematically as shown in Fig. 5b–f. A long waveinitially propagates from the upwelled fluid volume towardthe sloping topography (Fig. 5b). This wave reflects fromthe slope and travels back toward the vertical wall (Fig. 5c),reflecting once more and steepening as it travels. Upon re-turning toward the slope, a high-frequency wave packet hasformed with Etotal /PEo ø 15% (Fig. 5d). These waves breakupon the slope losing ;10% of the PEo to local turbulentdissipation and mixing (Fig. 5e).

The observed energy loss at the slope in Pusiano may becompared with the predicted energy loss using the internalIribarren model put forth in Fig. 10. The nondimensionalparameters j`, jsech, and Er/Ei were calculated from the fielddata on days 203 and 205. In Fig. 10b, the parameter j` isshown to be an accurate measure of Er/Ei in both the fieldand in the laboratory. Spilling and plunging breakers arepredicted, with the field measurements being grouped amongthe laboratory data collected in this study and that fromMichallet and Ivey (1999). Conversely, the parameter jsech isshown to overestimate the energy loss upon the slope inPusiano relative to the laboratory model in Fig. 10c. Inspec-tion of Fig. 14c and Table 2 reveals that the wind forcing ismoderate (i.e., W21 ø 0.4); consequently, the high-frequencywaves in Pusiano are sinusoidal in profile and thus will pos-sess a greater wave length than that given for a sech2 profileby Eq. (8b). In Fig. 10c, underestimation of the actual wavelength (and hence overestimation of the wave slope) willresult in a spurious decrease in jsech and corresponding over-estimation of Er/Ei. For these waves, jsin is more suitable.

An interpretation of the wave spectrum

A general model for the internal wave spectrum in lakeswas proposed by Imberger (1998). This model is analogousto the Garrett–Munk spectrum found in the oceanographicliterature. We use the results presented herein and the recentwork by Antenucci and Imberger (2001) and Boegman et al.(2003) to update this model for large lakes with a seasonalthermocline and a discernible frequency bandwidth betweenthe motions at the basin and buoyancy scales (bandwidth .102;103 Hz). The three main features of the internal wavespectrum are shown in Fig. 15a. First is the presence ofdiscrete natural and forced basin-scale seiches at frequenciesbetween zero and ;1024 Hz. Second is the presence of in-

stabilities, generated by the surface wind forcing and thebaroclinic shear of the basin-scale motions in the lake inte-rior, with frequencies ;1022 Hz. The third feature is theintermediate portion of the spectrum, which is shown to bedominated by freely propagating nonlinear wave groups ca-pable of breaking at the lake perimeter (depending on j).Moderate forcing (0.3 , W21 , 1), and perhaps topographyin rotational systems (Saggio and Imberger 1998; Wake etal. 2004), is required for excitation of these waves, whichare observed to have sinusoidal profiles at ;1024 Hz (Fig.14c) and sech2 profiles at ;1023 Hz (Saggio and Imberger1998; Boegman et al. 2003). It is interesting to note thatLake Kinneret, while strongly forced, does not appear togenerate a strong nonlinear internal wave response; yet thespectral peak resulting from shear instability near ;1022 Hzis enlarged relative to Biwa and Pusiano (Fig. 15a). Drawingon examples found in the literature, the behavior of the spec-trum in the nonlinear 1024 to 1023 Hz bandwidth may begeneralized according to the ratio of the wave height and thedepth of the surface layer a/h1 (or equivalently ho/h1 5 W21

because a ; ho). This ratio is commonly used to gauge thenonlinearity of progressive internal waves (e.g., Stanton andOstrovsky 1998). A consistent trend is found throughout ob-servations from a variety of natural systems with differingscales. Figure 15b shows sinusoidal waves near 1024 Hz andsech2 waves near 1023 Hz when a/h1 , 0.4 and a/h1 . 0.4,respectively. The waves associated with shear instability inLake Kinneret are mechanistically and observationally in-consistent with this model (a/h1 ø 0.2 and ;1022 Hz). Therevised spectral model for stratified lakes is summarized inTable 3.

Combination of the observations in Fig. 15b and the jmodel presented in Fig. 10 suggest that, in lakes with mod-erate to steep boundary slopes, 10% # R # 50% and 5%# G # 25%, in the Sulu Sea, R ; 40% and 15% # G #20% and, in the St. Lawrence estuary, R ; 6% to 8% (Table2). These calculations may differ by as much as 50% fromthose calculated using the ratio of the wave length to theslope length (e.g., Michallet and Ivey 1999; Bourgault andKelley 2003) and are only applicable to waves shoaling intoa closed slope at the depth of the pycnocline.

In stratified waterbodies, internal waves provide the cru-cial energy transfer between the large-scale motions forcedby winds and tides and the small-scale turbulent dissipationand mixing along sloping boundaries. This energy flux oc-curs through downscale spectral transfer and shoaling ofhigh-frequency internal waves along topography suited forwave breaking. In lakes, moderate wind forcing (0.3 , ho/h1 , 1) excites sub–basin-scale nonlinear wave groups thatpropagate throughout the basin. These wave groups are alsotidally generated in coastal regions. In both systems, when0.3 , ho/h1 ; a/h1 , 0.4, the waves have frequencies near1024 Hz and a sinusoidal profile whereas, when 0.4 , ho/h1

; a/h1 , 1, the waves have frequencies near 1023 Hz anda sech2 profile.

Using laboratory experiments, we have quantified thedown-scale spectral energy flux for closed basins with a two-layer stratification. High-frequency internal solitary wavesevolve from the basin-scale motions as t $ Ts and containup to 20% of the available potential energy introduced by

Page 16: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1635Internal waves in closed basins

Fig. 15. (a) Spectra of the vertically integrated potential energy signal (see Antenucci et al.2000) from Lake Pusiano (Fig. 14b) and lakes Kinneret and Biwa (see Boegman et al. 2003). Thedata were sampled at 10-s intervals and the spectra have been smoothed in the frequency domainto improve confidence at the 95% level, as shown by the dotted lines. Nmax denotes the maximumbuoyancy frequency. (b) Observations of wave nonlinearity, profile, and frequency from the liter-ature. Sources are as given in Table 2. For buoyancy period calculations, we have assumed N ;Nmax ; 1022 Hz.

Table 3. A revised spectral model for stratified lakes. Summa-rized from the data presented in Fig. 15.

WaveFrequency

(Hz)Buoyancy

periods

Linear basin scaleNonlinear sinusoidal profileNonlinear sech2 profileShear instabilityMaximum buoyancy frequency

,1024

;1024

;1023

;1022

;1022

.100;100;10;1;1

the wind. As Ti /2 is the characteristic wave travel time overL, the ratio Ti /Ts describes the wave evolution and representsa balance between high-frequency wave growth throughnonlinear steepening and high-frequency wave degenerationthrough shoaling at the boundary.

Wave breaking was observed at the boundary in the formof spilling, plunging, collapsing, and Kelvin–Helmholtzbreakers. A single wave breaking event resulted in 10% to75% of the incident wave energy being lost to dissipationand mixing, the remaining energy being found in a reflectedlong wave. The energy loss was dependent on the breakertype. Comparison with the results by Michallet and Ivey(1999) suggests that the mixing efficiency was also depen-dent on the breaker type and ranged between 5% and 25%.These boundary processes were modeled in terms of an in-ternal Iribarren number j, defined as the ratio of the wave

slope to the boundary slope. This definition is easily gen-eralized to waves shoaling into a closed slope at the depthof the pycnocline in lakes, oceans, and estuaries. For closedbasins, knowledge of the wave profile allowed j to be recastusing properties of the quiescent fluid and the forcing dy-namics.

Further study is required, as the laboratory facility did notpermit examination of extremely mild slopes, such as thosecommonly found in many lakes and coastal oceans. Theseresults will facilitate the parameterization of nonhydrostaticand sub–grid-scale wave processes into field-scale hydro-dynamic models (e.g., Boegman et al. 2004).

References

ANTENUCCI, J. P., AND J. IMBERGER. 2001. On internal waves nearthe high frequency limit in an enclosed basin. J. Geophys. Res.106: 22465–22474.

, J. IMBERGER, AND A. SAGGIO. 2000. Seasonal evolution ofthe basin-scale internal wave field in a large stratified lake.Limnol. Oceanogr. 45: 1621–1638.

APEL, J. R., J. R. HOLBROOK, A. K. LIU, AND J. J. TSAI. 1985. TheSulu Sea internal soliton experiment. J. Phys. Oceanogr. 15:1625–1650.

BATTIES, J. A. 1974. Surf similarity. Proc. Coastal Engineering Con-ference, ASCE, 14: 466–479.

BENNEY, D. J. 1966. Long nonlinear waves in fluid flows. J. Math.Phys. 45: 52–63.

BOEGMAN, L. 2004. The degeneration of internal waves in lakes

Page 17: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1636 Boegman et al.

with sloping topography. Ph.D. thesis, Centre for Water Re-search, The University of Western Australia.

, J. IMBERGER, G. N. IVEY, AND J. P. ANTENUCCI. 2003.High-frequency internal waves in large stratified lakes. Limnol.Oceanogr. 48: 895–919.

, G. N. IVEY, AND J. IMBERGER. 2004. An internal solitarywave parameterization for hydrodynamic lake models. Proc.15th Australasian Fluid Mechanics Conference.

, , AND . 2005. The energetics of large-scaleinternal wave degeneration in lakes. J. Fluid Mech. 531: 159–180.

BOURGAULT, D., AND D. E. KELLEY. 2003. Wave-induced mixingin a partially mixed estuary. J. Marine Res. 61: 553–576.

COWEN, E. A., K.-A. CHANG, AND Q. LIAO. 2001. A single-cameracoupled PTV-LIF technique. Exp. Fluids 31: 63–73.

DJORDJEVIC, V. D., AND L. G. REDEKOPPP. 1978. The fission anddisintegration of internal solitary waves moving over two-di-mensional topography. J. Phys. Oceanogr. 8: 1016–1024.

FARMER, D. M. 1978. Observations of long nonlinear internal wavesin a lake. J. Phys. Oceanogr. 8: 63–73.

, AND L. ARMI. 1999. Stratified flow over topography: Therole of small-scale entrainment and mixing in flow establish-ment. Proc. R. Soc. Lond. A 457: 3221–3258.

FERRIER, A. J., D. R. FUNK, AND P. J. W. ROBERTS. 1993. Appli-cation of optical techniques to the study of plumes in stratifiedfluids. Dyn. Atmos. Oceans 20: 155–183.

FISCHER, H. B., E. J. LIST, R. C. Y. KOH, J. IMBERGER, AND N. H.BROOKS. 1979. Mixing in inland and coastal waters. Academic.

FRICKER, P. D., AND H. M. NEPF. 2000. Buoyancy effects and in-ternal seiche decay. Proc. International Symposium on Strati-fied Flow 5: 313–318.

GALVIN, C. J. 1968. Breaker type classification on three laboratorybeaches. J. Geophys. Res. 73: 3651–3659.

GLOOR, M., A. WUEST, AND D. M. IMBODEN. 2000. Dynamics ofmixed bottom boundary layers and its implications for diapyc-nal transport in a stratified, natural water basin. J. Geophys.Res. 105: 8629–8646.

GOUDSMIT, G.-H., F. PETERS, M. GLOOR, AND A. WUEST. 1997.Boundary versus internal diapycnal mixing in stratified naturalwaters. J. Geophys. Res. 102: 27903–27914.

HALPERN, D. 1971. Observations of short-period internal waves inMassachusetts Bay. J. Marine Res. 29: 116–132.

HAMMACK, J. L., AND H. SEGUR. 1978. Modelling criteria for longwater waves. J. Fluid Mech. 84: 359–373.

HEAPS, N. S., AND A. E. RAMSBOTTOM. 1966. Wind effects on thewater in a narrow two-layered lake. Phil. Trans. R. Soc. Lond.A 259: 391–430.

HELFRICH, K. R. 1992. Internal solitary wave breaking and run-upon a uniform slope. J. Fluid Mech. 243: 133–154.

, AND W. K. MELVILLE. 1986. On nonlinear internal wavesover slope-shelf topography. J. Fluid Mech. 167: 285–308.

HODGES, B. R., J. IMBERGER, B. LAVAL, AND J. APPT. 2000. Mod-eling the hydrodynamics of stratified lakes. Proc. Hydroinfor-matics Conference 4: 23–27.

HOLLOWAY, P. E. 1987. Internal hydarulic jumps and solitons at ashealf break region on the Austrailan North West Shelf. J. Geo-phys. Res. 92: 5405–5416.

HORN, D. A., J. IMBERGER, AND G. N. IVEY. 1999. Internal solitarywaves in lakes—a closure problem for hydrostatic models.Proc. Aha Huliko Hawaiian Winter Workshop—Internal Grav-ity Waves II 11: 95–100.

, , AND . 2001. The degeneration of large-scale interfacial gravity waves in lakes. J. Fluid Mech. 434:181–207.

, L. G. REDEKOPP, J. IMBERGER, AND G. N. IVEY. 2000.

Internal wave evolution in a space-time varying field. J. FluidMech. 424: 279–301.

HUNKINS, K., AND M. FLIEGEL. 1973. Internal undular surges inSeneca Lake: A natural occurrence of solitons. J. Geophys.Res. 78: 539–548.

IMBERGER, J. 1998. Flux paths in a stratified lake: a review, p. 1–18. In J. Imberger [ed.], Physical processes in lakes and oceans.Coastal and Estaurine Studies. V. 54.

, AND G. N. IVEY. 1991. On the nature of turbulence in astratified fluid. Part 2: Application to lakes. J. Phys. Oceanogr.21: 659–680.

IVEY, G. N., AND J. IMBERGER. 1991. On the nature of turbulencein a stratified fluid. Part 1: The energetics of mixing. J. Phys.Oceanogr. 21: 650–658.

, AND R. I. NOKES. 1989. Vertical mixing due to the breakingof critical internal waves on sloping boundaries. J. Fluid Mech.204: 479–500.

KAKUTANI, T., AND N. YAMASAKI. 1978. Solitary waves on a two-layer fluid. J. Phys. Soc. Japan 45: 674–679.

KAO, T. W., F. S. PAN, AND D. RENOUARD. 1985. Internal solitonson the pycnocline generation, propagation, shoaling and break-ing over a slope. J. Fluid Mech. 159: 19–53.

KOMAR, P. D. 1976. Beach Processes and Sedimentation. Prentice-Hall.

LAMB, K. G. 2002. A numerical investigation of solitary internalwaves with trapped cores formed via shoaling. J. Fluid Mech.451: 109–144.

LEDWELL, J. R., AND B. M. HICKEY. 1995. Evidence for enhancedboundary mixing in the Santa Monica basin. J. Geophys. Res.100: 20665–20679.

LEE, C. Y., AND R. C. BEARDSLEY. 1974. The generation of longnonlinear internal waves in a weakly stratified shear flow. J.Geophys. Res. 79: 453–462.

LEGG, S., AND A. ADCROFT. 2003. Internal wave breaking at con-cave and convex continental slopes. J. Phys. Oceanogr. 33:2224–2246.

MACINTYRE, S., K. M. FLYNN, R. JELLISON, AND J. R. ROMERO.1999. Boundary mixing and nutrient fluxes in Mono Lake, Cal-ifornia. Limnol. Oceanogr. 4: 512–529.

MICHALLET, H., AND E. BARTHELEMY. 1997. Ultrasonic probes anddata processing to study interfacial solitary waves. Exp. Fluids22: 380–386.

, AND G. N. IVEY. 1999. Experiments on mixing due to in-ternal solitary waves breaking on uniform slopes. J. Geophys.Res. 104: 13467–13477.

MILES, J. W. 1981. The Korteweg–de Vries equation: A historicalessay. J. Fluid Mech. 106: 131–147.

MONISMITH, S. G. 1987. Modal response of reservoirs to windstress. ASCE J. Hydraul. Eng. 113: 1290–1306.

MORTIMER, C. H. 1974. Lake hydrodynamics. Mitt. Internat. Verein.Limnol. 20: 124–197.

, AND W. HORN. 1982. Internal wave dynamics and theirimplications for plankton biology in the Lake of Zurich. Mitt.Internat. Verein. Limnol. 127: 299–318.

MOUM, J. N., D. M. FARMER, W. D. SMYTH, L. ARMI, AND S. VA-GLE. 2003. Structure and generation of turbulence at interfacesstrained by internal solitary waves propagating shoreward overthe continental shelf. J. Phys. Oceanogr. 33: 2093–2112.

ORR, M. H., AND P. C. MIGNEREY. 2003. Nonlinear internal wavesin the South China Sea: Observation of the conversion of de-pression internal waves to elevation internal waves. J. Geo-phys. Res. 108: 3064. Doi:10.1029/2001JC001163.

OSTROVSKY, I., Y. Z. YACOBI, P. WALLINE, AND I. KALIKHMAN.1996. Seiche-induced mixing: Its impact on lake productivity.Limnol. Oceanogr. 4: 323–332.

Page 18: Boegman, L., G. N. Ivey, and J. Imberger. The …1620 Limnol. Oceanogr., 50(5), 2005, 1620–1637 q 2005, by the American Society of Limnology and Oceanography, Inc. The degeneration

1637Internal waves in closed basins

SAGGIO, A., AND J. IMBERGER. 1998. Internal wave weather in strat-ified lakes. Limnol. Oceanogr. 43: 1780–1795.

SANDSTROM, H., AND J. A. ELLIOTT. 1984. Internal tide and solitonson the Scotian Shelf: A nutrient pump at work. J. Geophys.Res. 89: 6415–6426.

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.

STANTON, T. P., AND L. A. OSTROVSKY. 1998. Observations of high-ly nonlinear internal solitons over the Continental Shelf. Geo-phys. Res. Lett. 25: 2695–2698.

STEVENS, C., AND J. IMBERGER. 1996. The initial response of astratified lake to a surface shear stress. J. Fluid Mech. 312: 39–66.

SVEEN, J. K., Y. GUO, P. A. DAVIES, AND J. GRUE. 2002. On thebreaking of internal solitary waves at a ridge. J. Fluid Mech.469: 161–188.

THOMPSON, R.O.R.Y., AND J. IMBERGER. 1980. Response of a nu-merical model of a stratified lake to wind stress. Proc. Intl.Symp. Stratified Flows 2: 562–570.

THORPE, S. A. 1971. Asymmetry of the internal seiche in LochNess. Nature 231: 306–308.

, A. HALL, AND I. CROFTS. 1972. The internal surge in LochNess. Nature 237: 96–98.

TORRENCE, C., AND G. P. COMPO. 1998. A practical guide to waveletanalysis. Bull. Amer. Meteor. Soc. 79: 61–78.

VLASENKO, V., AND K. HUTTER. 2002. Numerical experiments onthe breaking of solitary internal waves over a slope-shelf to-pography. J. Phys. Oceanogr. 32: 1779–1793.

WAKE, G. W., G. N. IVEY, AND J. IMBERGER. 2004. The temporalevolution of baroclinic basin-scale waves in a rotating circularbasin. J. Fluid Mech. 515: 63–86.

WALLACE, B. C., AND D. L. WILKINSON. 1988. Run-up of internalwaves on a gentle slope in a two-layered system. J. FluidMech. 191: 419–442.

WHITHAM, G. B. 1974. Linear and nonlinear waves. Wiley.WIEGAND, R. C., AND E. C. CARMACK. 1986. The climatology of

internal waves in a deep temperate lake. J. Geophys. Res. 91:3951–3958.

WUEST, A., G. PIEPKE, AND D. C. SENDEN. 2000. Turbulent kineticenergy balance as a tool for estimating vertical diffusivity inwind-forced stratified waters. Limnol. Oceanogr. 45: 1388–1400.

ZHOU, X., AND R. GRIMSHAW. 1989. The effect of variable currentson internal solitary waves. Dyn. Atmos. Oceans 14: 17–39.

ZIEGENBEIN, J. 1969. Short period internal waves in the Strait ofGibralter. Deep-Sea Res. 16: 479–487.

Received: 3 August 2004Accepted: 18 April 2005Amended: 30 May 2005