Skip to main content

Advertisement

Log in

Interpretation of coastal wind transfer functions with momentum balances derived from idealized numerical model simulations

  • Published:
Ocean Dynamics Aims and scope Submit manuscript

Abstract

The local wind-driven circulation off southern San Diego is addressed with two complementary statistical and dynamical frameworks based on observations and idealized numerical model simulations. The observations including surface currents from high-frequency radars, subsurface currents from a nearshore mooring, and wind records at a local wind station are analyzed with the idealized ocean model (MITgcm) simulations using realistic bottom topography and spatially uniform wind stress forcing. Statistically estimated anisotropic local wind transfer functions characterize the observed oceanic spectral response to wind stress separately in the x (east-west) and y (north-south) directions. We delineate the coastal circulation at three primary frequencies [low (σ L=0.0767 cycles per day (cpd)), diurnal (σ D=1 cpd), and inertial (σ f=1.07 cpd) frequencies] with the momentum budget equations and transfer functions. At low frequency, the magnitudes of transfer functions are enhanced near the coast, attributed to geostrophic balance between wind-driven pressure gradients and the Coriolis force on currents. The response diminishes away from the coast, returning to the balance between frictional and Coriolis terms, as in the classic Ekman model. On the contrary, transfer functions in the near-inertial frequency band show reduced magnitudes near the coast primarily due to friction, but exhibits the enhanced seaward response as a result of the inertial resonance. The idealized model simulations forced by local wind stress can identify the influences of remote wind stress and the biases in the data-derived transfer functions.

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

Similar content being viewed by others

References

  • Adcroft A, Campin J, Dutkiewicz S, Evangelinos C, Ferreira D, Forget G, Fox-Kemper B, Heimbach P, Hill C, Hill E (2014) MITgcm user manual. MIT Department of EAPS, Cambridge

    Google Scholar 

  • Allen JS (1980) Models of wind-driven currents on the continental shelf. Annu Rev Fluid Mech 12:389–433. doi:10.1146/annurev.fl.12.010180.002133

    Article  Google Scholar 

  • Allen JS, Smith RL (1981) On the dynamics of wind-driven shelf currents. Philos Trans Roy Soc London Ser A Math Phys Sci 302(1472):617–634. doi:10.1098/rsta.1981.0187

    Article  Google Scholar 

  • Allen J, Newberger P, Federiuk J (1995) Upwelling circulation on the Oregon continental shelf. Part I: Response to idealized forcing. J Phys Oceanogr 25:1843–1866

    Article  Google Scholar 

  • Austin J, Lentz S (2002) The inner shelf response to wind-driven upwelling and downwelling. J Phys Oceanogr 32(7):2171–2193

    Article  Google Scholar 

  • Brink KH (1991) Coastal-trapped waves and wind-driven currents over the continental shelf. Annu Rev Fluid Mech 23:389–412. doi:10.1146/annurev.fl.23.010191.002133

    Article  Google Scholar 

  • Brink K, Halpern D, Smith R (1980) Circulation in the Peruvian upwelling system near 15 S. J Geophys Res 85(C7):4036–4048

    Article  Google Scholar 

  • Castelao RM, Barth JA (2005) Coastal ocean response to summer upwelling favorable winds in a region of alongshore bottom topography variations off Oregon. J Geophys Res 110:C10S04. doi:10.1029/2004JC002409

    Google Scholar 

  • Chen C, Reid R, Nowlin W (1996) Near-inertial oscillations over the Texas-Louisiana shelf. J Geophys Res 101(C2):3509–3524. doi:10.1029/95JC03395

    Article  Google Scholar 

  • Craig PD (1988) Space and time scales in shelf circulation modelling. Cont Shelf Res 8(11):1221–1246

    Article  Google Scholar 

  • Craig PD (1989) Constant-eddy-viscosity models of vertical structure forced by periodic winds. Cont Shelf Res 9(4):343–358

    Article  Google Scholar 

  • Crawford GB, Large WG (1996) A numerical investigation of resonant inertial response of the ocean to wind forcing. J Phys Oceanogr 26(6):873–891

    Article  Google Scholar 

  • D’Asaro EA (1985) The energy flux from the wind to the near-inertial motions in the surface mixed layer. J Phys Oceanogr 15(8):1043–1059

    Article  Google Scholar 

  • Davies AM (2003) On the interaction between internal tides and wind-induced near-inertial currents at the shelf edge. J Geophys Res 108(C3):3099. doi:10.1029/2002JC001375

    Article  Google Scholar 

  • Davies A, Xing J (2003) Processes influencing wind-induced current profiles in near coastal stratified regions. Cont Shelf Res 23(14–15):1379–1400. doi:10.1016/S0278-4343(03)00119-5

    Article  Google Scholar 

  • Davis RE, Bogden PS (1989) Variability on the California shelf forced by local and remote winds during the Coastal Ocean Dynamics Experiment. J Geophys Res 94(C4):4763– 4783. doi:10.1029/JC094iC04p04763

    Article  Google Scholar 

  • Dever E (1997) Wind-forced cross-shelf circulation on the northern California shelf. J Phys Oceanogr 27(8):1566–1580

    Article  Google Scholar 

  • Edwards C, Fake T, Bogden P (2004a) Spring-summer frontogenesis at the mouth of Block Island Sound: 1. A numerical investigation into tidal and buoyancy-forced motion. J Geophys Res 109:C12021. doi:10.1029/2003JC002132

    Article  Google Scholar 

  • Edwards C, Fake T, Codiga D, Bogden P (2004b) Spring-summer frontogenesis at the mouth of Block Island Sound: 2. Combining acoustic Doppler current profiler records with a general circulation model to investigate the impact of subtidal forcing. J Geophys Res 109:C12022. doi:10.1029/2003JC002133

    Article  Google Scholar 

  • Ekman VW (1905) On the influence of the Earth’s rotation on ocean-currents. Ark Mat Astron Fys 2:1–53

    Google Scholar 

  • Elipot S, Gille ST (2009) Ekman layers in the Southern Ocean: spectral models and observations, vertical viscosity and boundary layer depth. Ocean Sci 5:115–139. http://www.ocean-sci.net/5/115/2009/

    Article  Google Scholar 

  • Emery WJ, Thomson RE (1997) Data analysis methods in physical oceanography. Elsevier, Boston

    Google Scholar 

  • Fewings M, Lentz S, Fredericks J (2008) Observations of cross-shelf flow driven by cross-shelf winds on the inner continental shelf. J Phys Oceanogr 38(11):2358–2378. doi:10.1175/2008JPO3990.1

    Article  Google Scholar 

  • Fukumori I, Lee T, Cheng B, Menemenlis D (2004) The origin, pathway, and destination of Nino-3 water estimated by a simulated passive tracer and its adjoint. J Phys Oceanogr 34:582–604

    Article  Google Scholar 

  • Gill AE (1982) Atmosphere-ocean dynamics, international geophysics series, vol. 30. Academic, San Diego

    Google Scholar 

  • Gonella J (1972) A rotary-component method for analysis in meteorological and oceanographic vector time series. Deep Sea Res 19:833–846

    Google Scholar 

  • Gopalakrishnan G, Cornuelle BD, Hoteit I (2013a) Adjoint sensitivity studies of Loop current and eddy shedding in the Gulf of Mexico. J Geophys Res 118(7):3315–3335. doi:10.1002/jgrc.20240

    Article  Google Scholar 

  • Gopalakrishnan G, Cornuelle BD, Hoteit I, Rudnick DL, Owens WB (2013b) State estimates and forecasts of the Loop current in the Gulf of Mexico using the MITgcm and its adjoint. J Geophys Res 118(7):3292–3314. doi:10.1002/jgrc.20239

    Article  Google Scholar 

  • Hodur RM (1997) The Naval Research Laboratory’s Coupled Ocean/Atmosphere Mesoscale Prediction System (COAMPS). Mon Wea Rev 125:1414–1430

    Article  Google Scholar 

  • Hoteit I, Cornuelle B, Kim SY, Forget G, Kohl A, Terrill E (2009) Assessing 4D-VAR for dynamical mapping of coastal high-frequency radar in San Diego. Dyn Atmos Oceans 48:175–197. doi:10.1016/j.dynatmoce.2008.11.005

    Article  Google Scholar 

  • Hoteit I, Cornuelle B, Heimbach P (2010) An eddy permitting variational data assimilation system for estimating the of the tropical Pacific. J Geophys Res 115:C03001. doi:10.1029/2009JC005347

    Google Scholar 

  • Hoteit I, Hoar T, Gopalakrishnan G , Collins N, Anderson J, Cornuelle B, Heimbach P (2013) A MITgcm/DART ensemble analysis and prediction system with application to the Gulf of Mexico. Dyn Atmos Oceans 63:1–12. doi:10.1016/j.dynatmoce.2013.03.002

    Article  Google Scholar 

  • Kim SY (2009) Coastal ocean studies in southern San Diego using high-frequency radar derived surface currents. Ph.D. thesis. Scripps Institution of Oceanography Technical Report. http://escholarship.org/uc/item/2z5660f4

  • Kim SY (2010) Observations of submesoscale eddies using high-frequency radar-derived kinematic and dynamic quantities. Cont Shelf Res 30:1639–1655. doi:10.1016/j.csr.2010.06.011

    Article  Google Scholar 

  • Kim SY (2014) A statistical description on the wind-coherent responses of sea surface heights off the U.S. West Coast. Ocean Dyn 64(1):29–46. doi:10.1007/s10236-013-0668-3

    Article  Google Scholar 

  • Kim SY, Terrill EJ, Cornuelle BD (2008) Mapping surface currents from HF radar radial velocity measurements using optimal interpolation. J Geophys Res 113:C10023. doi:10.1029/2007JC004244

    Article  Google Scholar 

  • Kim SY, Terrill EJ, Cornuelle BD (2009a) Assessing coastal plumes in a region of multiple discharges: The U.S.–Mexico border. Environ Sci Technol 43(19):7450–7457. doi:10.1021/es900775p

    Article  Google Scholar 

  • Kim SY, Cornuelle BD, Terrill EJ (2009b) Anisotropic response of surface currents to the wind in a coastal region. J Phys Oceanogr 39(6):1512–1533. doi:10.1175/2009JPO4013.1

    Article  Google Scholar 

  • Kim SY, Terrill EJ, Cornuelle BD, Jones B, Washburn L, Moline MA, Paduan JD, Garfield N, Largier JL, Crawford G, Kosro PM (2011) Mapping the U.S. West Coast surface circulation: A multiyear analysis of high-frequency radar observations. J Geophys Res 116:C03011. doi:10.1029/2010JC006669

    Google Scholar 

  • Kim SY, Cornuelle BD, Terrill EJ (2010a) Decomposing observations of high-frequency radar derived surface currents by their forcing mechanisms: Decomposition techniques and spatial structures of decomposed surface currents. J Geophys Res 115:C12007. doi:10.1029/2010JC006222

    Article  Google Scholar 

  • Kim SY, Cornuelle BD, Terrill EJ (2010b) Decomposing observations of high-frequency radar derived surface currents by their forcing mechanisms: Locally wind-driven surface currents. J Geophys Res 115:C12046. doi:10.1029/2010JC006223

    Article  Google Scholar 

  • Kim SY, Cornuelle BD, Terrill EJ, Jones B, Washburn L, Moline MA, Paduan JD, Garfield N, Largier JL, Crawford G, Kosro PM (2013) Poleward propagating subinertial alongshore surface currents off the U.S. West Coast. J Geophys Res 118(12):6791– 6806. doi:10.1002/jgrc.20400

    Article  Google Scholar 

  • Kim SY, Kosro PM, Kurapov AL (2014) Evaluation of directly wind-coherent near-inertial surface currents off Oregon using a statistical parameterization and analytical models. J Geophys Res 119. doi:10.1002/2014JC010115

  • Kirincich AR, Barth JA, Grantham BA, Menge BA, Lubchenco J (2005) Wind-driven inner-shelf circulation off central Oregon during summer. J Geophys Res 110:C10S03. doi:10.1029/2004JC002611

    Google Scholar 

  • Kohl ADS, Cornuelle B (2007) Interannual to decadal changes in the ECCO global synthesis. J Phys Oceanogr 37:313– 337. doi:10.1175/JPO3014.1

    Article  Google Scholar 

  • Large W, McWilliams J, Doney S (1994) Oceanic vertical mixing: A review and a model with a non-local boundary layer parameterization. Rev Geophys 32(4):363–403. doi:10.1029/94RG01872

    Article  Google Scholar 

  • Lee T, Williams E, Wang J, Evans R, Atkinson L (1989) Response of South Carolina continental shelf waters to wind and Gulf Stream forcing during winter of 1986. J Geophys Res 94(C8):10,715–10,754. doi:10.1029/JC094iC08p10715

    Article  Google Scholar 

  • Legg S, Hallberg R, Girton J (2006) Comparison of entrainment in overflows simulated by z-coordinate, isopycnal and non-hydrostatic models. Ocean Model 11(1–2):69–97. doi:10.1016/j.ocemod.2004.11.006

  • Lentz SJ (1995) Sensitivity of the inner-shelf circulation to the form of the eddy viscosity profile. J Phys Oceanogr 25(1):19– 28

    Article  Google Scholar 

  • Lentz SJ (2001) The influence of stratification on the wind-driven cross-shelf circulation over the North Carolina shelf. J Phys Oceanogr 31:2749–2760

    Article  Google Scholar 

  • Lentz SJ, Winant CD (1986) Subinertial currents on the southern California shelf. J Phys Oceanogr 16:1737– 1750

    Article  Google Scholar 

  • Lentz SJ, Trowbridge J (2001) A dynamical description of fall and winter mean current profiles over the northern California shelf. J Phys Oceanogr 31:914–931

    Article  Google Scholar 

  • Lentz SJ, Chapman DC (2004) The importance of nonlinear cross-shelf momentum flux during wind-driven coastal upwelling. J Phys Oceanogr 34:2444–2457 . doi:10.1175/JPO2644.1

    Article  Google Scholar 

  • Lentz S, Fewings M (2012) The wind-and wave-driven inner-shelf circulation. Annu Rev Mar Sci:317–343. doi:10.1146/annurevmarine-120709-142745

  • Lentz SJ, Guza R, Elgar S, Feddersen F, Herbers T (1999) Momentum balances on the North Carolina inner shelf. J Geophys Res 104:18,205–18,226. doi:10.1029/1999JC900101

    Article  Google Scholar 

  • Lentz SJ, Trowbridge J (2001) A dynamical description of fall and winter mean current profiles over the northern California shelf. J Phys Oceanogr 31:914–931

    Article  Google Scholar 

  • Lerczak JA, Hendershott MC, Winant CD (2001) Observations and modeling of coastal internal waves driven by a diurnal sea breeze. J. Geophys. Res. 106(C9):19,715–19,729

    Article  Google Scholar 

  • Lewis DM, Belcher SE (2004) Time-dependent, coupled, Ekman boundary layer solutions incorporating Stokes drift. Dyn Atmos Oceans 37:313–351. doi:10.1016/j.dynatmoce.2003.11.001

    Article  Google Scholar 

  • Madsen OS (1977) A realistic model of the wind-induced Ekman boundary layer. J Phys Oceanogr 7:248– 255

    Article  Google Scholar 

  • Marshall J, Adcroft A, Hill C, Perelman L, Heisey C (1997) A finite-volume, incompressible Navier Stokes model for studies of the ocean on parallel computers. J Geophys Res 102(C3):5753–5766. doi:10.1029/96JC02775

    Article  Google Scholar 

  • Menemenlis D, Fukumori I, Lee T (2005) Using Green’s functions to calibrate an ocean general circulation model. Mon Wea Rev 133(5):1224–1240. doi:10.1175/MWR2912.1

    Article  Google Scholar 

  • Millero FJ, Poisson A (1981) International one-atmosphere equation of state of seawater. Deep Sea Res 28A(6):625– 629

    Article  Google Scholar 

  • Millot C, Crepon M (1981) Inertial oscillations on the continental shelf of the Gulf of Lions: Observations and theory. J Phys Oceanogr 11:639–657

    Article  Google Scholar 

  • Oppenheim AV, Schafter RW, Buck JR (1998) Discrete-time signal processing, 2nd edn. Prentice Hall Signal Processing Series. Prentice-Hall, Inc., Upper Saddle River, NJ

    Google Scholar 

  • Pettigrew NR (1981) The dynamics and kinematics of the coastal boundary layer off Long Island. Ph.D. thesis, Massachusetts Institute of Technology. Dept. of Meteorology and Physical Oceanography. http://hdl.handle.net/1721.1/27913

  • Pollard RT, Millard RC (1970) Comparison between observed and simulated wind-generated inertial oscillations. Deep Sea Res 17:813–821

    Google Scholar 

  • Ponte ALS (2010) Periodic wind-driven circulation in an elongated and rotating basin. J Phys Oceanogr 40:2043–2058 . doi:10.1175/2010JPO4235.1

    Article  Google Scholar 

  • Price JF, Weller RA, Pinkel R (1986) Diurnal cycling: Observations and models of the upper ocean response to diurnal heating, cooling, and wind mixing. J Geophys Res 91(C7):8411– 8427. doi:10.1029/JC091iC07p08411

  • Pringle JM (2002) Enhancement of wind-driven upwelling and downwelling by alongshore bathymetric variability. J Phys Oceanogr 32:3101–3112

    Article  Google Scholar 

  • Rio MH, Hernandez F (2003) High-frequency response of wind-driven currents measured by drifting buoy and altimetry over the World Ocean. J Geophys Res 108(C8):3238 . doi:10.1029/2002JC001655

    Google Scholar 

  • Rykaczewski R, Checkley D (2008) Influence of ocean winds on the pelagic ecosystem in upwelling regions. Proc Natl Acad Sci 105(6):1965–1970. doi:10.1073/pnas.0711777105

    Article  Google Scholar 

  • Shay LK, Lee TN, Williams EJ, Graber HC, Rooth CGH (1998) Effects of low-frequency current variability on near-inertial submesoscale vorticies. J. Geophys. Res. 103(C9):18,691–18,714. doi:10.1029/98JC01007

  • Shearman RK (2005) Observations of near-inertial current variability on the New England shelf. J Geophys Res 110:C02012. doi:10.1029/2004JC002341

    Google Scholar 

  • Shen CY, Evans TE (2001) Surface-to-subsurface velocity projection for shallow water currents. J Geophys Res 106(C4):6973–6984. doi:10.1029/2000JC000267

  • Stammer D, Wunsch C, Giering R, Eckert C, Heimbach P, Marotzke J, Adcroft A, Hill C, Marshall J (2002) Global ocean circulation during 1992–1997, estimated from ocean observations and a general circulation model. J Geophys Res 107(C9):3118. doi:10.1029/2001JC000888

    Article  Google Scholar 

  • von Storch H, Zwiers F (1999) Statistical analysis in climate research. Cambridge University Press

  • Todd R, Rudnick D, Mazloff M, Davis R, Cornuelle B (2011) Poleward flows in the southern California Current System: Glider observations and numerical simulation. J Geophys Res 116:C02026. doi:10.1029/2010JC006536

  • Verdy A, Mazloff M, Cornuelle BD, Kim SY (2013) Wind-driven sea level variability on the california coast: An adjoint sensitivity analysis. J Phys Oceanogr 44(1):297–318. doi:10.1175/JPO-D-13-018.1

  • Weisberg RH, Li Z, Muller-Karger R (2001) West Florida shelf response to local wind forcing: April 1998. J Geophys Res 106(C12):31,239–31,262. doi:10.1029/2000JC000529

    Article  Google Scholar 

  • Weller RA (1981) Observations of the velocity response to wind forcing in the upper ocean. J Geophys Res 86(C3):1969–1977. doi:10.1029/JC086iC03p01969

    Article  Google Scholar 

  • Winant CD (2004) Three-dimensional wind-driven flow in an elongated rotating basin. J Phys Oceanogr 34:462–476

    Article  Google Scholar 

  • Winant CD, Beardsley RC, Davis RE (1987) Moored wind, temperature, and current observations made during Coastal Ocean Dynamics Experiments 1 and 2 over the Northern California continental shelf and upper slope. J Geophys Res 92(C2):1569–1604. doi:10.1029/JC092iC02p01569

    Article  Google Scholar 

  • Wunsch C (1996) The ocean circulation inverse problem. Cambridge University Press, NY

    Book  Google Scholar 

  • Xing J, Davies A (2004) On the influence of a surface coastal front on near-inertial wind-induced internal wave generation. J Geophys Res 109(C1):C01023 . doi:10.1029/2003JC001794

    Google Scholar 

  • Yelland M, Taylor PK (1996) Wind stress measurements from the open ocean. J Phys Oceanogr 26:541–558

    Article  Google Scholar 

  • Young IR (1999) Wind generated ocean waves, vol. 2. Elsevier, London

    Google Scholar 

  • Zhang X, Cornuelle B, Roemmich D (2011) Adjoint sensitivity of the Nino-3 surface temperature to wind forcing~. J Clim 24:4480– 4493. doi:10.1175/2011JCLI3917.1

    Article  Google Scholar 

Download references

Acknowledgments

Sung Yong Kim is supported by the Basic Science Research Program through the National Research Foundation (NRF), Ministry of Education (no. 2013R1A1A2057849), the Human Resources Development of the Korea Institute of Energy Technology Evaluation and Planning (KETEP), Ministry of Trade, Industry and Energy (no. 20134030200300), and the program on Management of Marine Organisms Causing Ecological Disturbance and Harmful Effects through Korea Integrated Marine Science Technology (KIMST) and Ministry of Fishery (MOF), Republic of Korea. Ganesh Gopalakrishnan and Aurelien Ponte are sponsored by NOAA (NOAA Award NA10OAR4320156). Surface current data are provided from the Southern California Coastal Ocean Observing System (SCCOOS, Available online at http://www.sccoos.org) at Scripps Institution of Oceanography. Wind data at the Tijuana River (Tidal Linkage) are maintained by the System-Wide Monitoring Program at the Tijuana River National Estuarine Research Reserve, Estuarine Reserves Division (TRNERR), NOAA, Centralized Data Management Office (CDMO), Baruch Marine Field Lab, and University of South Carolina. The bathymetry is provided by National Geophysical Data Center (NGDC). The authors thank B. D. Cornuelle for providing ideas and a chance to generate synergy among coauthors and E. Terrill for providing the data.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sung Yong Kim.

Additional information

Responsible Editor: Pierre Lermusiaux

Appendices

Appendix A: Spatial scales of the surface wind field

The decorrelation length scales of the surface wind field off southern San Diego are examined with the COAMPS nowcast data with two spatial resolutions of 1.7 and 5.1 km (Hodur 1997). An example of spatial correlations of the hourly model wind at W in Fig. 1 for one year (2006) is shown in Fig 13. The length scales are asymmetric, i.e., relatively long seaward and short shoreward, in the range of 25 to 120 km.

Fig. 13
figure 13

An example ofspatial correlations of the COAMPS hourly nowcast wind data at W (a white asterisk) in Fig. 1. a and c: ρ u u (x,x ). b and d: ρ v v (x,x ). a and b: 1.7 km resolution. b and d: 5.1 km resolution

Appendix B: Equilibrium of the model solutions

An example of model currents forced by low-frequency cross-shore wind stress (τ x ) is presented with the hourly time series of u- and v-components at the center of a cross-shore line a in Fig. 1 and the surface layer (z=−0.5 m) over 30 days (Fig. 14a and b). The data for the first three days are excluded (a black box), and the rest of model solutions is used for the transfer function analysis and momentum budget analysis. Although the weak near-inertial oscillations are still imposed in the time series after three days, the model solutions at low frequency reach the equilibrium quickly.

Fig. 14
figure 14

The hourly time series of model currents, forced by low-frequency wind stress in the cross-shore diction (τ x ), at the center of a cross-shore line a in Fig. 1 and the surface layer (z=−0.5 m) over 30 days. The data for he first three days are excluded (a black box), and the rest of model solutions is used for the transfer function analysis. a u-component (cm s−1), b v-component (cm s−1)

Appendix C: Isotropic subsurface transfer function

From the isotropic relationship (4), the wind-coherent currents are as follows:

$$ \hat{\mathbf{u}}_{\mathrm{W}}(z, \sigma) = \mathbf{\mathsf{E}}(z,\sigma) \left[\hat{\uptau}_{x}(\sigma) + i\hat{\uptau}_{y}(\sigma)\right], $$
(24)
$$ {\kern3.1pc} = \left[ {\mathsf{E}}_{r}(z, \sigma) + i {\mathsf{E}}_{i}(z, \sigma)\right] \left[\hat{\uptau}_{x}(\sigma) + i\hat{\uptau}_{y}(\sigma)\right], $$
(25)

where E r and E i indicate the real and imaginary parts of the isotropic transfer function (E). The corresponding magnitude and argument are referred to as E and ΘE, respectively, i.e.,

$$ {\kern.7pc} E = |\mathbf{\mathsf{E}}| = |\mathsf{E}_{r} + i\mathsf{E}_{i}|, $$
(26)
$$ {\Theta}^{E} = \tan^{-1} \frac{\mathsf{E}_{i}}{\mathsf{E}_{r}}, $$
(27)

The data-derived isotropic transfer function (E) is computed from the TJR wind stress and the detided HFR surface and ADCP subsurface currents (Fig. 15a and c). The model-derived isotropic transfer function (F) is based on the Ekman model (Fig. 15b and d) (e.g., Kim et al. 2009).

Fig. 15
figure 15

Magnitudes and arguments of surface and subsurface isotropic transfer functions derived from HFRs and ADCP observations at T in Fig. 1 and the Ekman model. a E, b F, c ΘE, and d ΘF. f c and δ E denote the local inertial frequency (1.07 cpd) and the Ekman depth (\(\delta _{E} = \pi \sqrt {2\nu /f_{\mathrm {c}}} = 8.705\) m, where ν=3×10−4 m 2 s −1 is the kinematic viscosity), respectively

The peak near the inertial frequency decreases with depth, consistent with the model, penetrating up to about 8 m. However, the significant variance at low frequency does not appear in the model, which indicates the influence of pressure gradients near the coast because the Ekman model does not include pressure gradient terms (Fig. 15a and b) (e.g., Kim et al. 2009). The arguments of data- and model-derived transfer functions are very similar except for the sign shift at the zero frequency (Fig. 15c and d).

Appendix D: Time lag and veering angle

As the argument (ΘE) of the isotropic transfer function (E in Eq. 24) contains the time lag (α) and veering angle (β), i.e., ΘE=α+β (see Kim et al. (2009)),

$$ \hat{u}(\sigma) + i\hat{v}(\sigma) = E(\sigma)e^{i {\Theta}^{E}(\sigma)}\left[\hat{\uptau}_{x}(\sigma) + i\hat{\uptau}_{y}(\sigma)\right], $$
(28)

we can separate them with the relationship between individual components of wind stress and currents, i.e., anisotropic transfer functions.

As a simple case when ISCs are dominant over ASCs (e.g., offshore region), the anisotropic transfer functions can be paired (H x x =H y y and H x y = −H y x ) (e.g., Kim et al. 2009). The Fourier coefficients of currents are expressed with the isotropic convention:

$$ \hat{u} + i\hat{v} = \left(H_{xx}\hat{\uptau}_{x} + H_{xy}\hat{\uptau}_{y} \right) + i\left(H_{yx}\hat{\uptau}_{x} + H_{yy}\hat{\uptau}_{y}\right), $$
(29)
$$\begin{array}{@{}rcl@{}} {\kern2.2pc} &=& \left(|H_{xx}|e^{i{\Theta}_{xx}^{H}}\hat{\uptau}_{x} + |H_{xy}| e^{i{\Theta}_{xy}^{H}}\hat{\uptau}_{y}\right)\\ &&+ i\left(|H_{xy}|e^{i\left({\Theta}_{xy}^{H}+\pi\right)}\hat{\uptau}_{x} + |H_{xx}| e^{i{\Theta}_{xx}^{H}}\hat{\uptau}_{y}\right), \end{array} $$
(30)
$$ {\kern2.2pc} = \left(|H_{xx}|e^{i{\Theta}_{xx}^{H}} -i |H_{xy}|e^{i{\Theta}_{xy}^{H}}\right)\left(\hat{\uptau}_{x} + i\hat{\uptau}_{y}\right), $$
(31)

where H x x has the zero veering angle and the time lag of \({\Theta }_{xx}^{H}\) (\(\alpha = {\Theta }_{xx}^{H}\) and β=0), and H x y has the veering angle of −90 and the time lag of \({\Theta }_{xy}^{H}\) (\(\alpha = {\Theta }_{xy}^{H}\) and β=−90). Thus, this comparison enables us to differentiate the veering angle and time lag when the argument is ambiguously computed under the isotropic assumption such as complex coherence and correlation.

At \(\sigma = \sigma _{\mathrm {L}}, {\Theta }_{xx}^{H}\) is in the range of −2 to −5, and \({\Theta }_{xy}^{H}\) is in the range of −7 to −8 (Fig. 2m–p; Eq. 31). The estimated time lag (α) and veering angle (β) for three cases of the argument (ΘE) at low frequency (ΘE=−30,−45, and −60) (e.g., Kim et al. 2010b) are shown in Table 1. For the case of σ=0 cpd (Ekman model), the time lang and veering angle are zero and −45, respectively.

Table 1 Estimates of the time lag (α, degrees or hours) and veering angle (β, degrees) for three cases of the argument (ΘE=α+β) of the isotropic transfer function at low frequency (σ=σ L) using the pairs of time lags (\({\Theta }_{xx}^{H}\) or \({\Theta }_{xy}^{H}\)) and veering angles (0 or −90) of the anisotropic transfer functions. The negative time lag indicates that wind stress leads the currents, and the negative veering angle measures the amount of turn that the currents head to the right of the wind direction

At σ=σ D, \({\Theta }_{xx}^{H}\) is in the range of 2 to 6, and \({\Theta }_{xy}^{H}\) is in the range of −61 to −67 (Fig. 7m–p). The estimated time lag (α) and veering angle (β) for four cases of the argument (ΘE) at the diurnal frequency (ΘE=−30, −45,−60, and −75) (e.g., Kim et al. 2010b) are shown in Table 2.

Table 2 Estimates of the time lag (α, degrees or hours) and veering angle (β, degrees) for four cases of the argument (ΘE=α+β) of the isotropic transfer function at the diurnal frequency (σ=σ D) using the pairs of time lags (\({\Theta }_{xx}^{H}\) or \({\Theta }_{xy}^{H}\)) and veering angles (0 or −90) of the anisotropic transfer functions. The meaning of the time lag and the convention of the veering angle are noted in Table 1

Appendix E: Estimates of ISCs and ASCs

Equations 14 and 15 are solved together frequency by frequency (e.g., Wunsch 1996; Kim 2009):

$$\begin{array}{@{}rcl@{}} \left[\begin{array}{l} \hat{u}\\ \hat{v}\\ \end{array}\right] & = \left[\begin{array}{llll} \hat{\tau}_{x} & \hat{\tau}_{y} & \hat{\tau}_{x} & \hat{\tau}_{y} \\ \hat{\tau}_{y} & -\hat{\tau}_{x} & -\hat{\tau}_{y} & \hat{\tau}_{x}\\ \end{array}\right] \left[\begin{array}{l} \mathsf{I}_{d}\\ \mathsf{I}_{c}\\ \mathsf{A}_{d}\\ \mathsf{A}_{c}\\ \end{array}\right], \end{array} $$
(32)

where \(\hat {u}, \hat {u}, \hat {\uptau }_{x}\), and \(\hat {\uptau }_{y}\) are column vectors consisting of N sub-sampled Fourier coefficients at a single frequency ( is the vector transpose), presented as follows:

$$ {\kern4pt} \hat{u} = \left[ \hat{u}_{1} \hat{u}_{2}\cdots \hat{u}_{N}\right]^{\dag},\\ $$
(33)
$$ {\kern4pt} \hat{v} = \left[ \hat{v}_{1} \hat{v}_{2}\cdots \hat{v}_{N}\right]^{\dag},\\ $$
(34)
$$ \hat{\uptau}_{x} = \left[ \hat{\uptau}_{x1} \hat{\uptau}_{x2}\cdots \hat{\uptau}_{xN}\right]^{\dag},\\ $$
(35)
$$ \hat{\uptau}_{y} = \left[ \hat{\uptau}_{y1} \hat{\uptau}_{y2}\cdots \hat{\uptau}_{yN}\right]^{\dag}. $$
(36)

Equation 1 is formulated as

$$ \mathbf{d} = \mathbf{Z}\:\mathbf{m}, $$
(37)

where d,Z, and m correspond to the currents, wind stress, and isotropy and anisotropy components. The estimated model coefficients (\(\mathbf {\widehat {m}}\)) are

$$ \mathbf{\widehat{m}} = \mathbf{P}\mathbf{Z}^{\dag}\left(\mathbf{Z}\mathbf{P}\mathbf{Z}^{\dag} + \mathbf{R}\right)^{-1}\mathbf{d}, \\ $$
(38)
$$ {\kern.7pc} = \left(\mathbf{Z}^{\dag}\mathbf{R}^{-1}\mathbf{Z}+\mathbf{P}^{-1}\right)^{-1}\mathbf{Z}^{\dag}\mathbf{R}^{-1}\mathbf{d}, $$
(39)

where P and R are the model and error covariance matrices ( is the matrix transpose).

When the error covariance matrix (R) is replaced with an identity matrix (I), the model covariance matrix (P) will become a unique factor to adjust the inverse of snr (R P −1; β k , k= 1, 2, 3, and 4), whose diagonal components reflect the contribution of each basis function:

$$\begin{array}{@{}rcl@{}} \mathbf{P} = \left[\begin{array}{llll} \beta_{1} & 0 &0 & 0\\ 0 & \beta_{2} & 0 & 0\\ 0 & 0 & \beta_{3}& 0 \\ 0 & 0 & 0 &\beta_{4}\\ \end{array}\right]. \end{array} $$
(40)

The inverse of snr for the estimates of ISCs and ASCs is assumed as 0.1, 0.1, 0.5, and 0.5 of the mean eigenvalues of the covariance matrix Fourier coefficients of wind stress, respectively, for I d, I c, A d, and A c at the given three frequencies (σ = σ L, σ D, and σ f). The chosen ratios are less sensitive to the overall results.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kim, S.Y., Gopalakrishnan, G. & Ponte, A. Interpretation of coastal wind transfer functions with momentum balances derived from idealized numerical model simulations. Ocean Dynamics 65, 115–141 (2015). https://doi.org/10.1007/s10236-014-0766-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10236-014-0766-x

Keywords

Navigation