Skip to main content
Log in

The baroclinic response to wind in a small two-basin lake

  • Research Article
  • Published:
Aquatic Sciences Aims and scope Submit manuscript

Abstract

Field data and two linear layered models were used to examine the baroclinic response to wind in a small elongated two-basin lake, Amisk Lake (Alberta, Canada). For the first vertical baroclinic mode, wind-forced horizontal modes were simulated using a dynamic two-layer variable cross-section (TVC) model. The first horizontal mode, H1, was found to dominate the exchange between the two basins of Amisk Lake. The highest velocities associated with H1 occurred in the constricted channel connecting the two basins. This high velocity led to strong damping which brought H1 in near-resonance with the diurnal wind despite a difference in periods. Along with H1, the second horizontal mode, H2, was detected at a thermistor mooring. The TVC showed that H2 resulted from the coupling between along-thalweg wind distribution and the bathymetry of Amisk Lake. The damping for H2 was found to be weaker than for H1, likely because of weaker bottom drag in the connecting channel. The evolution of vertical H1 modes were simulated using a multi-layered box model. In response to wind pulses, the first vertical mode V1H1 initially dominated over higher vertical modes causing two-layer exchange. Following the faster decay of V1H1, higher vertical modes shifted the exchange to three and more layers. Our study shows the importance of the coupling between wind stress distribution and lake bathymetry in exciting horizontal modes, reveals the effect of damping on resonance with wind, and explains the evolution of exchange associated with vertical modes.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

Notes

  1. Note that r = 0 represents the barotropic mode.

References

  • Antenucci JP, Imberger J (2003) The seasonal evolution of wind/internal wave resonance in Lake Kinneret. Limnol Oceanogr 48(5):2055–2061

    Article  Google Scholar 

  • Antenucci JP, Imberger J, Saggio A (2000) Seasonal evolution of the basin-scale internal wave field in a large stratified lake. Limnol Oceanogr 45(7):1621–1638

    Article  Google Scholar 

  • Arneborg L, Liljebladh B (2001) The internal seiches in Gullmar Fjord. Part I: dynamics. J Phys Oceanogr 31(9):2549–2566

    Article  Google Scholar 

  • Bäuerle E (1994) Transverse baroclinic oscillations in Lake überlingen. Aquat Sci 56(2):145–160

    Article  Google Scholar 

  • Bendat JS, Piersol AG (2010) Random data: analysis and measurement procedures. Wiley, New York

  • Bengtsson L (1973) Mathematical models of wind induced circulation in a lake. In: Hydrology of lakes, no. 9 in IAHS publication, IAHS Press, Gentbrugge, pp 313–320

  • Boegman L, Ivey GN (2007) Experiments on internal wave resonance in periodically forced lakes. In: 5th international symposium on environmental hydraulics, Tempe

  • Boegman L, Ivey GN, Imberger J (2005) The energetics of large-scale internal wave degeneration in lakes. J Fluid Mech 531:159–180

    Article  Google Scholar 

  • Boehrer B, Imberger J, Münnich KO (2000) Vertical structure of currents in western Lake Constance. J Geophys Res 105(C12):28,823–828,835

    Google Scholar 

  • Charnock H (1955) Wind stress on a water surface. Q J R Meteorol Soc 81:639–640

    Article  Google Scholar 

  • Chen CTA, Millero FJ (1986) Precise thermodynamic properties for natural waters covering only the limnological range. Limnol Oceanogr 31(3):657–662

    Article  CAS  Google Scholar 

  • De Silva CW (2007) Vibration: fundamentals and practice. CRC press, Boca Raton

  • Evans MA, MacIntyre S, Kling GW (2008) Internal wave effects on photosynthesis: experiments, theory, and modeling. Limnol Oceanogr 53:339–353

    Article  CAS  Google Scholar 

  • Fischer HB, List EJ, Koh RCY, Imberger J, Brooks NH (1979) Mixing in inland and coastal waters. Academic Press, New York

  • Gill AE (1982) Atmosphere-ocean dynamics, International Geophysics Series, vol 30. Academic Press, London

  • Gomez-Giraldo A, Imberger J, Antenucci JP (2006) Spatial structure of the dominant basin-scale internal waves in Lake Kinneret. Limnol Oceanogr 51(1):229–246

    Article  Google Scholar 

  • Heaps NS, Ramsbottom AE (1966) Wind effects on the water in a narrow two-layered lake. Philos Trans R Soc Lond Ser A 259(1102):391–430

    Article  Google Scholar 

  • Hutter K, Salvadé G, Schwab DJ (1983) On internal wave dynamics in the northern basin of the Lake of Lugano. Geophys Astrophys Fluid Dyn 27(3):299–336

    Article  Google Scholar 

  • Imboden DM, Lemmin U, Joller T, Schurter M (1983) Mixing processes in lakes: mechanisms and ecological relevance. Aquat Sci 45(1):11–44

    Article  CAS  Google Scholar 

  • Kenney BC (1979) Lake surface fluctuations and the mass flow through the narrows of Lake Winnipeg. J Geophys Res 84(C3):1225–1235

    Article  Google Scholar 

  • La Rocca M, Sciortino G, Boniforti M (2002) Interfacial gravity waves in a two-fluid system. Fluid Dyn Res 30(1):31–66

    Article  Google Scholar 

  • Lawrence GA, Burke JM, Murphy TP, Prepas EE (1997) Exchange of water and oxygen between the two basins of Amisk Lake. Can J Fish Aquat Sci 54(9):2121–2132

    CAS  Google Scholar 

  • Lemckert C, Antenucci J, Saggio A, Imberger J (2004) Physical properties of turbulent benthic boundary layers generated by internal waves. J Hydraul Eng 130:58–69

    Article  Google Scholar 

  • Lemmin U (1987) The structure and dynamics of internal waves in Baldeggersee. Limnol Oceanogr 32(1):43–61

    Google Scholar 

  • Lemmin U, Mortimer CH, Bäuerle E (2005) Internal seiche dynamics in Lake Geneva. Limnol Oceanogr 50(1):207–216

    Google Scholar 

  • Lorrai C, Umlauf L, Becherer JK, Lorke A, Wüest A (2011) Boundary mixing in lakes: 2. Combined effects of shear- and convectively induced turbulence on basin-scale mixing. J Geophys Res 116:C10018

    Google Scholar 

  • MacIntyre S, Melack JM (1995) Vertical and horizontal transport in lakes: linking littoral, benthic, and pelagic habitats. J N Am Benthol Soc 4(4):599–615

    Article  Google Scholar 

  • Monismith SG (1985) Wind-forced motions in stratified lakes and their effect on mixed-layer shear. Limnol Oceanogr 30(4):771–783

    Article  Google Scholar 

  • Monismith SG (1986) An experimental study of the upwelling response of stratified reservoirs to surface shear stress. J Fluid Mech 171:407–439

    Article  Google Scholar 

  • Mortimer CH (1952) Water movements in lakes during summer stratification; evidence from the distribution of temperature in Windermere. Philos Trans R Soc Lond Ser B 236(635):355–398

    Article  Google Scholar 

  • Mortimer CH (1953) The resonant response of stratified lakes to wind. Aquat Sci 15(1):94–151

    Article  Google Scholar 

  • Mortimer CH (1965) Spectra of long surface waves and tides in Lake Michigan and at Green Bay, Wisconsin. University of Michigan Great Lakes Research Division Publication 13

  • Mortimer CH (1979) Strategies for coupling data collection and analysis with dynamic modelling of lake motions. In: Hydrodynamics of lakes: proceedings of a symposium, Lausanne, pp 183–222

  • Münnich M, Wüest A, Imboden DM (1992) Observations of the second vertical mode of the internal seiche in an alpine lake. Limnol Oceanogr 37(8):1705–1719

    Article  Google Scholar 

  • Okely P, Imberger J (2007) Horizontal transport induced by upwelling in a canyon-shaped reservoir. Hydrobiologia 586(1):343–355

    Article  Google Scholar 

  • Ostrovsky I, Yacobi YZ, Walline P, Kalikhman I (1996) Seiche-induced mixing: Its impact on lake productivity. Limnol Oceanogr 41(2):323–332

    Article  Google Scholar 

  • Pannard A, Beisner BE, Bird DF, Braun J, Planas D, Bormans M (2011) Recurrent internal waves in a small lake: potential ecological consequences for metalimnetic phytoplankton populations. Limnol Oceanogr Fluids Environ 1:91–109

    Article  Google Scholar 

  • Platzman GW (1978) Normal modes of the world ocean. Part I. Design of a finite-element barotropic model. J Phys Oceanogr 8:323–343

    Article  Google Scholar 

  • Prepas EE (1990) Amisk Lake. In: Mitchell P, Prepas EE (eds) Atlas of Alberta lakes, University of Alberta Press, Edmonton, pp 225–231

  • Roget E, Salvadé G, Zamboni F (1997) Internal seiche climatology in a small lake where transversal and second vertical modes are usually observed. Limnol Oceanogr 42(4):663–673

    Article  Google Scholar 

  • Rueda F, Singleton V, Stewart M, Little J, Lawrence GA (2010) Modeling the fate of oxygen artificially injected in the hypolimnion of a lake with multiple basins: Amisk Lake revisited. In: Proceedings of the 6th international symposium on environmental hydraulics (ISEH), Athens

  • Saggio A, Imberger J (1998) Internal wave weather in a stratified lake. Limnol Oceanogr 43(8):1780–1795

    Google Scholar 

  • Shimizu K, Imberger J (2008) Energetics and damping of basin-scale internal waves in a strongly stratified lake. Limnol Oceanogr 53(4):1574–1588

    Article  Google Scholar 

  • Shimizu K, Imberger J, Kumagai M (2007) Horizontal structure and excitation of primary motions in a strongly stratified lake. Limnol Oceanogr 52(6):2641–2655

    Article  Google Scholar 

  • Spigel RH, Imberger J (1980) The classification of mixed-layer dynamics of lakes of small to medium size. J Phys Oceanogr 10(7):1104–1121

    Article  Google Scholar 

  • Thorpe SA (1974) Near-resonant forcing in a shallow two-layer fluid: a model for the internal surge in Loch Ness? J Fluid Mech 63(3):509–527

    Article  Google Scholar 

  • Umlauf L, Lemmin U (2005) Interbasin exchange and mixing in the hypolimnion of a large lake: the role of long internal waves. Limnol Oceanogr 50(5):1601–1611

    Article  Google Scholar 

  • Vidal J, Casamitjana X (2008) Forced resonant oscillations as a response to periodic winds in a stratified reservoir. J Hydraul Eng 134(4):416–425

    Article  Google Scholar 

  • Vidal J, Rueda FJ, Casamitjana X (2007) The seasonal evolution of high vertical-mode internal waves in a deep reservoir. Limnol Oceanogr 52(6):2656–2667

    Article  Google Scholar 

  • Wake GW, Hopfinger EJ, Ivey GN (2007) Experimental study on resonantly forced interfacial waves in a stratified circular cylindrical basin. J Fluid Mech 582:203–222

    Article  Google Scholar 

  • Wiegand RC, Carmack EC (1986) The climatology of internal waves in a deep temperate lake. J Geophys Res 91(C3):3951–3958

    Article  Google Scholar 

  • Wiegand RC, Chamberlain V (1987) Internal waves of the second vertical mode in a stratified lake. Limnol Oceanogr 32(1):29–42

    Article  Google Scholar 

Download references

Acknowledgments

Dr. Roger Pieters provided valuable feedback on an earlier draft of this manuscript. Digital bathymetry of Amisk Lake was acquired from the Alberta Geological Survey.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Y. E. Imam.

Additional information

Y. E. Imam was partially supported by NSERC Discovery Grants held by B. Laval and G. Lawrence.

Appendix: Model derivation

Appendix: Model derivation

The objective of this appendix is to give the necessary background for the derivation of the TVC and MLM models. We start with the governing equations for the response to wind in elongated lakes of variable cross-section and N density layers (“Governing equations and assumptions”). The equations are simplified and then limited to two layers for the derivation of the TVC (“Two-layer variable-cross-section model”). The governing equations for N layers are then restricted to uniform width and depth for derivation of the MLM model (“Multi-layer box model”).

Governing equations and assumptions

For a narrow elongated lake with N density layers, the linearized equations of momentum and mass conservation are given by,

$$ \frac{\partial q_{j}}{\partial t}=-gS_{j}\sum_{i=1}^{j}\left(\frac{\rho_{i}- \left(1-\delta_{1i}\right)\rho_{i-1}} {\rho_{j}}\right)\frac{\partial\eta_{i}}{\partial x}+ \frac{1}{\rho_{j}}\left(\delta_{1j}F_{s}-F_{d,j}\right) $$
(15)
$$ \frac{\partial q_{j}}{\partial x}+b_{j}\frac{\partial\eta_{j}}{\partial t}-\left(1-\delta_{Nj}\right)b_{j+1}\frac{\partial\eta_{j+1}}{\partial t}=0 $$
(16)

where δ1i , δ1j , and δ Nj are Kronecker delta, the index \(j=1,2\ldots N\) denotes the density layers and interfaces from the surface downwards, q j (xt) is the flow in layer j, η j (xt) is the deflection of interface jS j (x) and ρ j are the cross-sectional area and density of layer jb j (x) is the cross-sectional width at interface jF s (xt) is the wind forcing acting on the surface of the cross-section, and F d,j (xt) is the drag force on layer j. It should be noted that, for arriving at Eq. 15, hydrostatic pressure, no-rotation, and negligible cross-thalweg motions were assumed.

Further, by using a Guldberg–Mohn parameterization for the drag τ d,j  = 2κρ j q j as done in many similar studies (e.g., Heaps and Ramsbottom 1966; Bengtsson 1973; Platzman 1978; Bäuerle 1994; Shimizu and Imberger 2008), Eq. 15 simplifies to,

$$ \frac{\partial q_{j}}{\partial t}+2\kappa q_{j}=-gS_{j}\sum_{i=1}^{j}\left(\frac{\rho_{i}- \left(1-\delta_{1i}\right)\rho_{i-1}} {\rho_{j}}\right)\frac{\partial\eta_{i}}{\partial x}+\delta_{1j}\frac{F_{s}}{\rho_{j}} $$
(17)

where κ(xt) is the drag coefficient.

To eliminate η i from the above equation, differentiate it with respect to t

$$ \frac{\partial^{2}q_{j}}{\partial t^{2}}+2\frac{\partial\kappa q_{j}}{\partial t}=-gS_{j}\sum_{i=1}^{j}\left(\frac{\rho_{i}-\left(1-\delta_{1i}\right) \rho_{i-1}}{\rho_{j}}\right)\frac{\partial^{2}\eta_{i}}{\partial t\partial x}+ \delta_{1j}\frac{1}{\rho_{j}}\frac{\partial F_{s}}{\partial t}, $$
(18)

recast Eq. 16 into the form,

$$ \frac{1}{b_{j}}\sum_{i=j}^{N}\frac{\partial q_{i}}{\partial x}+\frac{\partial\eta_{j}}{\partial t}=0, $$
(19)

differentiate this form with respect to x, and use the result to eliminate ∂2η i /∂xt from Eq. 18. These steps give the governing equation for layer flow,

$$ \frac{\partial^{2}q_{j}}{\partial t^{2}}+2\frac{\partial\kappa q_{j}}{\partial t}=gS_{j}\sum_{i=1}^{j} \left(\frac{\rho_{i}-\left(1-\delta_{1i}\right)\rho_{i-1}} {\rho_{j}}\right)\frac{\partial}{\partial x} \left(\frac{1}{b_{i}}\sum_{j=i}^{N}\frac{\partial q_{j}}{\partial x}\right)+\delta_{1j}\frac{1}{\rho_{j}}\frac{\partial F_{s}}{\partial t} $$
(20)

Two-layer variable-cross-section model

For two-layer stratification \(\left(j=1,2\right),\) Eq. 20 can be recast as,

$$ \frac{\partial^{2}}{\partial t^{2}}\left(q_{2}-\frac{S_{2}}{S_{1}}\left(\frac{\rho_{1}} {\rho_{2}}\right)q_{1}\right)+2\frac{\partial}{\partial t}\kappa\left(q_{2}-\frac{S_{2}} {S_{1}}\left(\frac{\rho_{1}}{\rho_{2}}\right)q_{1}\right)-gS_{2}\left(\frac{\rho_{2}- \rho_{1}}{\rho_{2}}\right)\frac{\partial}{\partial x}\left(\frac{1}{b_{2}} \frac{\partial q_{2}}{\partial x}\right)=-\frac{1}{\rho_{2}}\frac{S_{2}}{S_{1}} \frac{\partial F_{s}}{\partial t}, $$
(21)

which can be simplified to Eq. 2 by using the rigid-lid (q 1 =  −q 2) and Boussinesq approximations (Gill 1982).

To solve Eq. 2, we first consider free modal evolution with no damping and assume separable layer flow with sinusoidal temporal dependence, \(q_{2}=Q_{n}(x)\sin\left(\omega_{u,n}t\right). \) For this case, Eq. 2 reduces to the eigenvalue problem described by Eq. 3 whose solution yields Q n and the undamped modal frequency ω u,n .

Next, consider wind forcing of arbitrary spatial distribution and periodic temporal dependence as given by,

$$ F_{s}=\sum_{n=1}^{\infty}F_{s}^{(n,m)}=-\left(\frac{S_{1}+S_{2}}{S_{2}} \right)\left(\sum_{n=1}^{\infty}A_{n}Q_{n}\right)\left(B_{m}\exp \left(\imath\omega_{f,m} t\right)\right). $$
(22)

Let the layer flow forced by this wind be given by the superposition of modal responses, \(q_{2}=\Upsigma_{n=1}^{\infty}M_{n,m}(t)Q_{n}(x), \) use a constant drag coefficient κ n for each mode such that \(\Upsigma_{n=1}^{\infty}\kappa_{n}M_{n,m}Q_{n}=\kappa(x,t)q_{2}, \) and substitute for F s q 2, and κq 2 in Eq. 2 to get,

$$ \sum_{n=1}^{\infty}Q_{n}\left\{ \frac{d^{2}M_{n,m}}{dt^{2}}+2\kappa_{n}\frac{dM_{n,m}}{dt}+ \omega_{n,r}^{2}M_{n,m} \right\} =\frac{1}{\rho_{1}}\sum_{n=1}^{\infty}A_{n}Q_{n}B_{m}\frac{d}{dt}\left\{ \exp\left(\imath\omega_{f,m}t\right)\right\} . $$
(23)

To derive an evolution equation for each mode (i.e., decouple the modes), multiply both sides of the eigenvalue problem (Eq. 3) by (S −11  + S −12 ) and note that the operators on both sides are self-adjoint which leads to the orthogonality property \(\int_{0}^{L}(S_{1}^{-1}+S_{2}^{-1})Q_{p}Q_{n}dx=0\) for pn. Multiply both sides of Eq. 23 by (S −11  + S −12 )Q p , integrate over x, and use the orthogonality property to obtain the ordinary differential equation (ODE) for a single mode,

$$ \frac{d^{2}M_{n,m}}{dt^{2}}+2\zeta_{n}\omega_{u,n} \frac{dM_{n,m}}{dt}+\omega_{u,n}^{2} M_{n,m}=\frac{\imath\omega_{f,m}}{\rho_{1}}A_{n}B_{m} \exp\left(\imath\omega_{f,m}t \right) $$
(24)

which describes the forced response of a damped linear oscillator having a damping ratio \(\zeta_{n}=\kappa_{n}\omega_{u,n}^{-1}.\) The solution to this ODE and the derivation of the TVC proceeds as outlined in Eqs. 58 and related text.

It only remains to note that, in Eq. 24, A n is obtained by decomposing the spatial wind-forcing distribution b 1(x) in terms of Q n

$$ b_{1}(x)=-\sum_{n=1}^{\infty}\left(\frac{S_{1}+S_{2}}{S_{2}}\right) (A_{n}Q_{n}). $$
(25)

This decomposition is done by multiplying both sides of Eq. 25 by Q p /S 1, integrating over the thalweg, and using the orthogonality property \(\int_{0}^{L}(S_{1}^{-1}+S_{2}^{-1})Q_{p}Q_{n}dx=0\) for pn to get,

$$ A_{n}=-\frac{\int_{0}^{L}S_{1}^{-1}Q_{n}b_{1}(x)dx}{\int_{0}^{L} (S_{1}^{-1}+S_{2}^{-1})Q_{n}^{2}dx}. $$
(26)

Multi-layer box model

For a basin having uniform width and depth, the multi-layer box model (MLM) simulates the response of individual VrHn modes to wind forcing. To simulate the individual responses, the MLM separates vertical modes (Vr) following the model of Monismith (1985) and separates horizontal modes (Hn) following the model of Heaps and Ramsbottom (1966). To derive the MLM, we start from Eq. 20 which governs the flow q j for N layers. By setting the width of the basin to a uniform width B and S j  = Bh j , Eq. 20 reduces to,

$$ \frac{\partial^{2}q_{j}}{\partial t^{2}}+2\frac{\partial\kappa q_{j}}{\partial t}-\sum_{k=1}^{N}\alpha_{jk}\frac{\partial^{2}q_{k}}{\partial x^{2}}=\delta_{1j}\frac{1}{\rho_{j}}\frac{\partial F_{s}}{\partial t} $$
(27)

where h j is the static-equilibrium thickness of layer j and the matrix α is a function of h j as defined earlier. Following the procedure outlined above for the TVC, Eq. 27 is reduced to the eigenvalue problem described by Eq. 10 by considering the case of undamped unforced motion and using \(q_{j}=R_{j}\sin(n\pi xL^{-1})\sin(\pi\beta tL^{-1}).\) The solution to this eigenvalue problem, β2 R = α R, yields the wave celerity β r and eigenvector R (r) for a vertical mode r. Now, consider the wind forcing given by \(F_{s}=A_{n}(t)\sin(n\pi xL^{-1})\) and assume a solution of the form \(q_{j}=\Upsigma_{r=0}^{\infty}R_{j}^{(r)}M_{n,r}(t)\sin(n\pi xL^{-1}).\) Substituting these expressions for F s and q j in Eq. 27 reduces it to,

$$ \sum_{r=0}^{N-1}\left\{ R_{j}^{(r)} \left(\frac{d^{2}M_{n,r}}{dt^{2}}+2\kappa_{n}\frac{dM_{n,r}}{dt}+ \omega_{n,r}^{2}M_{n,r}\right)\right\} =\frac{\delta_{1j}}{\rho_{j}}\frac{dA_{n}}{dt}. $$
(28)

By multiplying both sides by R (p) j ρ j h −1 j (where p denotes a different vertical mode) and summing over j, we get

$$ \frac{d^{2}M_{n,r}}{dt^{2}}+2\kappa_{n} \frac{dM_{n,r}}{dt}+\omega_{n,r}^{2}M_{n,r}= \frac{R_{1}^{(r)}h_{1}^{-1}}{\sum_{j=1}^{N} (R_{j}^{(r)})^{2}\rho_{j}h_{j}^{-1}}\frac{dA_{n}}{dt} $$
(29)

where the orthogonality property ∑ N j=1 R (p) j ρ j h j −1 R (r) j  = δ pr N j=1 (R (r) j ) 2ρ j h −1 j was used. This property can be proved by multiplying Eq. 10 by a diagonal matrix ϕ whose diagonal elements are ϕ jj  = ρ j h −1 j and noting that the matrices ϕ and ϕ−1α are real and symmetric which necessitates that, for the distinct eigenvalues β r , the corresponding eigenmodes are orthogonal.

For steady wind initiated at t = 0, the solution to Eq. 29 is as given earlier (Eq. 13) and can be used to obtain the solution for general wind conditions A n (t) using the convolution (Heaps and Ramsbottom 1966),

$$ M_{n,r}=\int\limits_{0}^{t}M_{n,r}^{{\rm st}}(t-t^{\prime})\frac{dA_{n}} {dt^{\prime}}dt^{\prime} $$
(30)

where M st n,r is the solution to steady wind (Eq. 13) and A n is given by (Heaps and Ramsbottom 1966),

$$ A_{n}(t)=\frac{\int_{0}^{L}\sin(n\pi xL^{-1})F_{s} (x,t)dx}{\int_{0}^{L}\sin^{2}(n\pi xL^{-1})dx}. $$
(31)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Imam, Y.E., Laval, B.E. & Lawrence, G.A. The baroclinic response to wind in a small two-basin lake. Aquat Sci 75, 213–233 (2013). https://doi.org/10.1007/s00027-012-0268-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00027-012-0268-1

Keywords

Navigation