Skip to main content

Advertisement

Log in

Extreme bed shear stress during coastal downwelling

  • Published:
Ocean Dynamics Aims and scope Submit manuscript

Abstract

The wind-driven circulation of coastal oceans has been studied for many decades. Using a 2.5-dimensional hydrodynamic model, this work unravels new aspects inherent with this circulation. In agreement with previous studies, downwelling-favorable coastal winds create an overturning cross-shelf circulation that operates to mix nearshore water. On timescales of days, this circulation tends to eliminate itself causing a “shutdown” of the cross-shelf circulation. For the first time, here, the author demonstrates that this shutdown is accompanied by creation of a zone of extremely high bed shear stresses (> 0.35 Pa) that operates to “plow” the seabed over an offshore distance of ~ 10–20 km. The author postulates that the associated sediment erosion episodes and their likely ammonification of the water column are key in the understanding of the biogeochemistry shaping coastal marine ecosystems.

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.

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

References

  • Allen JS, Newberger PA (1996) Downwelling circulation on the Oregon continental shelf. Part I: response to idealized forcing. J Phys Oceanogr 26:2011–2035

    Article  Google Scholar 

  • Allen JS, Newberger PA, 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 JA, Lentz SJ (2002) The inner shelf response to winddriven upwelling and downwelling. J Phys Oceanogr 22:2171–2193

    Article  Google Scholar 

  • Baines PG, Mitsudera H (1994) On the mechanism of shear flow instabilities. J Fluid Mech 276:327–342

    Article  Google Scholar 

  • Barton ED, Torres R, Figueiras FG, Gilcoto M, Largier J (2016) Surface water subduction during a downwelling event in a semi-enclosed bay. J Geophys Res-Oceans 121:7088–7107. https://doi.org/10.1002/2016JC011950

    Article  Google Scholar 

  • Burchard H, Petersen O, Rippeth TP (1998) Comparing the performance of the Mellor–Yamada and the k-ε two-equation turbulence models. J Geophys Res 103:10543–10554

    Article  Google Scholar 

  • Campin J-M, Hill C, Jones H, Marshall J (2011) Super-parameterization in ocean modeling: application to deep convection. Ocean Mod 36(1–2):90–101. https://doi.org/10.1016/j.ocemod.2010.10.003

    Article  Google Scholar 

  • Chang GC, Dickey TD, Williams AJ III (2001) Sediment resuspension over a continental shelf during hurricanes Edouard and Hortense. J Geophys Res 106:9517–9531

    Article  Google Scholar 

  • Churchill JH, Wirick CD, Flagg CN, Pietrafesa LJ (1994) Sediment resuspension over the continental shelf east of the Delmarva Peninsula. Deep-Sea Res II Top Stud Oceanogr 41:341–363

    Article  Google Scholar 

  • Cushman-Roisin B, Beckers J-M (2011) Introduction to geophysical fluid dynamics. Academic Press, Cambridge

    Google Scholar 

  • de Szoeke RA, Richman R (1984) On wind-driven mixed layers with strong horizontal gradients—a theory with application to coastal upwelling. J Phys Oceanogr 14:364–377

    Article  Google Scholar 

  • Drazin PG, Reid WH (1981) Hydrodynamic stability. Cambridge University Press, Cambridge

    Google Scholar 

  • Ekman VW (1905) On the influence of the earth’s rotation on ocean-currents. Ark Mat Astr Fys 2(11):1–53

    Google Scholar 

  • Eppley RW, Peterson BJ (1979) Particulate organic flux and planktonic new production in the deep ocean. Nature 282:677–680

    Article  Google Scholar 

  • Fanning KA, Carder KL, Betzer PR (1982) Sediment resuspension by coastal waters: a potential for nutrient re-cycling on the oceans margins. Deep Sea Res 29:953–965

    Article  Google Scholar 

  • Fennel W (1999) Theory of the Benguela upwelling system. J Phys Oceanogr 29(2):177–190

    Article  Google Scholar 

  • Greenwood B, Osborne PD (1990) Vertical and horizontal structure in cross-shore flows: an example of undertow and wave set-up on a barred beach. Coast Eng 14:543–580

    Article  Google Scholar 

  • Hansen LS, Blackburn TH (1992) Effect of algal bloom deposition on sediment respiration rates and fluxes. Mar Biol 112:147–152

    Article  Google Scholar 

  • Hela I (1976) Vertical velocity of the upwelling in the sea. Soc Sci Fennicia Commentationes Phys Math 46:9–24

    Google Scholar 

  • Herbert RA (1999) Nitrogen cycling in coastal marine ecosystems. FEMS Microbial Rev 23:563–590

    Article  Google Scholar 

  • Ings DW, Gregory RS, Schneider DC (2008) Episodic downwelling predicts recruitment of Atlantic cod, Greenland cod and white hake to Newfoundland coastal waters. J Mar Res 66(4):529–561. https://doi.org/10.1357/002224008787157476

    Article  Google Scholar 

  • Jensen HM, Lomstein E, Sorensen J (1990) Benthic NH4 and NO3 fluxes following sedimentation of a spring phytoplankton bloom in Aarhus bight, Denmark. Mar Ecol Prog Ser 61:87–96

    Article  Google Scholar 

  • Kämpf J (2009) Ocean modelling for beginners. Springer, Heidelberg

    Book  Google Scholar 

  • Kämpf J (2010) Advanced ocean modelling. Springer, Heidelberg

    Book  Google Scholar 

  • Kämpf J (2015) Interference of wind-driven and pressure gradient-driven flows in shallow homogeneous water bodies. Ocean Dyn 65(11):1399–1410. https://doi.org/10.1007/s10236-015-0882-2

    Article  Google Scholar 

  • Kämpf J (2017) Wind-driven overturning, mixing and upwelling in shallow water: a nonhydrostatic modeling study. J Mar Sci Eng 5:47. https://doi.org/10.3390/jmse5040047

    Article  Google Scholar 

  • Kämpf J, Backhaus JO (1998) Shallow, brine-driven free convection in polar oceans: nonhydrostatic numerical process studies. J Geophys Res 103:5577–5593

    Article  Google Scholar 

  • Kämpf J, Chapman P (2016) Upwelling systems of the world: a scientific journal to the most productive marine ecosystems. Springer International Publishing, Cham

    Book  Google Scholar 

  • Kämpf J, Fohrmann H (2000) Sediment-driven downslope flow in submarine canyons and channels: three-dimensional numerical experiments. J Phys Oceanogr 30(9):2302–2319

    Article  Google Scholar 

  • Kämpf J, Myrow PM (2018) Wave-created mud suspensions: a theoretical study. J Mar Sci Eng 6(2):29. https://doi.org/10.3390/jmse6020029

    Article  Google Scholar 

  • 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. https://doi.org/10.1029/2004JC002611

    Article  Google Scholar 

  • Kochergin VP (1987) Three-dimensional prognostic models. In: Heaps NS (ed) Three-dimensional coastal ocean models, coastal estuarine science series 4. American Geophysical Union, Washington, pp 201–208

    Chapter  Google Scholar 

  • Kuzmin D, Mierka O, Turek S (2007) On the implementation of the k-ε turbulence model in incompressible flow solvers based on a finite element discretization. Int J Comput Sci Math 1(2–4):193–206

    Article  Google Scholar 

  • Large WG, McWilliams JC, Doney SC (1994) Oceanic vertical mixing: a review and a model with a nonlocal boundary layer parameterization. Rev Geophys 32:363–403

    Article  Google Scholar 

  • Laws EA (2004) New production in the equatorial Pacific: a comparison of field data with estimates derived from empirical and theoretical models. Deep-Sea Res I 51:205–211

    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 

  • Lettmann KA, Wolff J-O, Badewien TH (2009) Modeling the impact of wind and waves on suspended particulate matter fluxes in the east Frisian Wadden Sea (southern North Sea). Ocean Dyn 59(2):239–262. https://doi.org/10.1007/s10236-009-0194-5

    Article  Google Scholar 

  • McCreary J, Kundu PK (1985) Western boundary circulation driven by an alongshore wind: with application to the Somali current system. J Mar Res 43:493–516

    Article  Google Scholar 

  • Mitchener H, Torfs H, Whitehouse R (1996) Erosion of mud/sand mixtures. Coast Eng 29:1–25

    Article  Google Scholar 

  • Moum JN, Perlin A, Klymak JM, Levine MD, Boyd T, Kosro PM (2004) Convectively-driven mixing in the bottom boundary layer over the continental shelf during downwelling. J Phys Oceanogr 34:2189–2202. https://doi.org/10.1175/1520-0485(2004)034<2189:CDMITB>2.0.CO;2

    Article  Google Scholar 

  • Nixon SW (1981) Remineralisation and nutrient cycling in coastal marine ecosystems. In: Nielson BJ, Cronin LE (eds) Estuaries and nutrients. Humana Press, New Jersey, pp 111–138

    Chapter  Google Scholar 

  • Rodi W (1987) Examples of calculation methods for flow and mixing in stratified fluids. J Geophys Res 92(C5):5305–5328

    Article  Google Scholar 

  • Schloen J, Stanev EV, Grashorn S (2017) Wave-current interactions in the southern North Sea: the impact on salinity. Ocean Model 111:19–37

    Article  Google Scholar 

  • Shanks AL, Brink L (2005) Upwelling, downwelling, and cross-shelf transport of bivalve larvae: test of a hypothesis. Mar Ecol Prog Ser 302:1–12

    Article  Google Scholar 

  • Stanev EV, Dobrynin M, Pleskachevsky A, Grayek S, Günther H (2009) Bed shear stress in the southern North Sea as an important driver for suspended sediment dynamics. Ocean Dyn 59(2):183–194. https://doi.org/10.1007/s10236-008-0171-4

    Article  Google Scholar 

  • Tilburg CE (2003) Across-shelf transport on a continental shelf: do across-shelf winds matter? J Phys Oceanogr 33:2675–2688

    Article  Google Scholar 

  • Wijesekera HW, Allen JS, Newberger PA (2003) Modeling study of turbulent mixing over the continental shelf: comparison of turbulent closure schemes. J Geophys Res 108(C3):3103. https://doi.org/10.1029/2001JC001234

    Article  Google Scholar 

Download references

Acknowledgments

The author thanks two referees for their fruitful comments and suggestions that have improved the quality of this work. All hydrodynamic model codes used in this work can be obtained from the author on request (jochen.kaempf@flinders.edu.au).

Conflict of interest

The author declares that he has no competing interests.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jochen Kämpf.

Additional information

Responsible Editor: Dirk Olbers

Appendices

Simplified Turbulence Closure Scheme

1.1 Description

This additional study applies the same shelf model as in the main study, but with one modification. Instead of using the k-ε scheme, vertical eddy diffusivity, Az, is diagnosed here from Kochergin’s turbulence closure (Kochergin 1987) that can be written as

$$ {A}_z={\left(c\Delta z\right)}^2\sqrt{{\left(\partial u/\partial z\right)}^2+{\left(\partial v/\partial z\right)}^2-{N}^2} $$
(A1)

where the free parameter is set to c = 0.2. Vertical eddy diffusivity, Kz, is again based on a turbulent Prandtl number of 0.7. The lower bound of Az is set to a molecular value of 10−6 m2/s. The upper bound of Az is set to 0.1 m2/s. A value of Az = 0.05 m2/s is applied near the sea surface as a representation of background wind stirring of surface water.

Note that Az in (A1) becomes negligibly small as the gradient Richardson number (19) approaches unity. Hence, additional treatment is required to parameterize dynamic instabilities such as convective or Kelvin-Helmholtz instabilities that are expected to occur when Rig < 1/4 (e.g., Baines and Mitsudera 1994; Cushman-Roisin and Beckers 2011). To account for this, the scheme is amended by the additional parameterization of effective eddy viscosity as a function of the local gradient Richardson number (e.g., Large et al. 1994; Wijesekera et al. 2003); that is;

$$ {A}_z=\left\{\begin{array}{c}{A}_{\ast }{\left[1-{\left({\mathrm{Ri}}_{\mathrm{g}}/{\mathrm{Ri}}_{cr}\right)}^2\right]}^3\ \mathrm{when}\kern0.5em 0<{\mathrm{Ri}}_{\mathrm{g}}<{\mathrm{Ri}}_{cr}\ \\ {}{A}_{\ast}\kern6.5em \mathrm{when}\kern0.5em {\mathrm{Ri}}_{\mathrm{g}}<0\end{array}\right.\kern0.5em $$
(A2)

Here, we use the theoretical value of Ricr = ¼ in conjunction with A* = 0.1 m2/s in experiment E-1 and a reduced value of A* = 0.01 m2/s in experiment E-2. Whenever Rig < Ricr the value from (A2) is always used to override that from (A1). Hereby it should be firstly noted that the choice of A* = 0.1 m2/s follows from features of the mixing zone in the control experiment (see Fig. 6) and it is much larger than typically used to simulate turbulent diffusivity/viscosity in the interior of the water column (A* ~ 0.005 m2/s). Secondly, it should also be explained that purpose of reducing A* to 0.01 m2/s in the second experiment is to demonstrate that the high bed shear stresses developing near the downwelling front are the consequence of enhanced vertical momentum diffusion. Otherwise the experiments E-1 and E-2 are identical to the configuration of the control experiment (D-1, see Table 1).

1.2 Results

Experiment E-1 largely reproduces the results of the control experiment except for an overestimation of turbulent stirring near vertical boundaries (Appendix Fig. 14, compared with Fig. 6). This result is not unexpected given that, unlike in the k-ε scheme, the mixing length is not limited in vicinity of boundaries here. As a consequence both Ekman layers are thicker than in the control experiment. Nevertheless, all other scales are remarkable close to the control prediction, including the progression of bed shear stresses (Appendix Fig. 15, compare with Fig. 8). Again, the shelf model predicts the offshore progression of a peak bed shear stress slightly above 0.4 Pa that moves offshore at a rate of ~ 3 km per day. While experiment E-2 also predicts a similar progression due to the offshore displacement of the downwelling jet, the resultant maximum bed shear stresses are significantly smaller (Appendix Fig. 15), as least on time scales < 5 days. Hence, it is the shear instability process (induced by the cross-shelf circulation) that significantly enhanced bed shear stresses via vertical mixing of the along-shelf momentum. Quod erat demonstrandum.

High-Resolution Nonhydrostatic Simulations

1.1 Description

The shelf model used in this work cannot resolve the special scales of nonhydrostatic turbulent vortices inherent with the Kelvin Helmholtz instability mechanism which have an aspect ratio (ratio between horizontal and vertical scales) of unity. In order to resolve those scales, the hydrodynamic equations detailed in Section 2.1.3 are applied here with a finer grid spacing of Δx = Δz = 1 m on a smaller horizontal spatial scale.

The nonhydrostatic model is initialized by the predictions from the control experiment (D-1) using variable values for a selected single grid column (which has a horizontal width of 1 km) at a selected time. Hence, the nonhydrostatic model considers a 1-km wide model domain of a total depth corresponding to the location of the control experiment. This one-way coupling is done after every day of the “mother” simulation and at an interval of 2 km (i.e., for very second grid column of the shelf model) to an offshore distance of 30 km. Altogether this gives 10 × 15 = 150 simulations, but it is sufficient to only present the results of a few selected results. The total simulation time of nonhydrostatic model runs is 4 h using a numerical time step of Δt = 1 s. Initially small random fluctuations are added to the density field to seed minuscule fluctuations that can grow as part of instability processes. The vertical velocity field starts with zero values.

The nonhydrostatic assumes a flat seafloor (on the spatial scale of 1 km), and uses the same wind-stress forcing as the shelf model. Additionally the nonhydrostatic model also accounts for an external barotropic pressure-gradient force (due to the sloping surface), also prescribed from the shelf model. This pressure-gradient force is kept constant over the simulation period (5 h). Furthermore, the model uses cyclic horizontal boundaries, which ignores any lateral advection effects. The model also adopts Kochergin’s turbulence closure but without parameterization of shear-flow and convective instabilities, which are resolved in the model. For isotropic turbulence, Kochergin’s turbulence scheme can be expressed by (e.g., Kämpf and Backhaus 1998):

$$ {A}_x={A}_z={c}^2{\left(\Delta x\Delta z\right)}^2\sqrt{{\left(\partial u/\partial z\right)}^2+{\left(\partial v/\partial z\right)}^2+{\left(\partial v/\partial x\right)}^2+{\left(\partial w/\partial x\right)}^2-{N}^2} $$
(B1)

where c = 0.2. In addition, a turbulent Prandt number of unity is assumed (Kx = Kz = Ax = Az). The aim of this supplementary study is to test whether the vertical shear of u in conjunction with the density stratification predicted by the shelf model can initiate Kelvin-Helmholtz instabilities and, if so, whether, this mechanism leads to the predicted enhancement of bed shear stresses via modulation of v. To illustrate any stirring mechanism, the nonhydrostatic model also predicts the evolution of a passive concentration field from the advection-diffusion equation:

$$ \frac{\partial }{\partial t}C+u\frac{\partial }{\partial x}C+w\frac{\partial }{\partial z}C=\frac{\partial }{\partial x}\left({K}_x\frac{\partial }{\partial x}C\right)+\frac{\partial }{\partial z}\left({K}_z\frac{\partial }{\partial z}C\right) $$
(B2)

which is of the same form as the density conservation Eq. (4). Initially, C varies linearly between zero and unity over the depth of the water column, using zero-flux vertical and cyclic lateral boundary conditions.

1.2 Results

Only results for the simulations that start from day 5 of the mother simulation are discussed here. Other start times yielded similar results. Within the stirring zone, for instance at xo = 12 km (see Fig. 6), the nonhydrostatic model predicts the onset of shear-flow instabilities within 2–3 h of simulation (Appendix Fig. 16). The instabilities start to develop first near the seafloor before filling the entire water column. In contrast, outside the mixing zone at xo = 30 km, the pronounced density stratification near the bottom of the water column (see Fig. 6) prevents turbulence generation in vicinity of the seafloor (Appendix Fig. 17). Hence, the bed shear stress remains at moderate levels.

Appendix Fig. 18 displays the evolution of bed shear stresses at xo = 12 km. Initially of bed shear stress rapidly decreases uniformly in the entire mode domain over the first hour of simulation. This decrease is caused by a modified vertical eddy viscosity that leads to a decrease of the along-shelf velocity component v near the seafloor. After the onset of dynamic instabilities, which is apparent from the increase of the standard deviation of bed shear stresses, the maximum bed shear stresses “bounce back” to reach almost the same value (~ 0.4 Pa) as simulated by the shelf model.

Simulation for other offshore locations confirm that the entire zone of apparently high vertical eddy viscosity/diffusivity (see Fig. 6d) is prone to the onset of shear-flow instabilities and vigorous mixing in the entire water column (Appendix Fig. 18). The nonhydrostatic simulations also confirm that it is the downward mixing of long-shelf momentum v that substantially enhance bed shear stresses to extremely high values.

Fig. 14
figure 14

Same as Fig. 6, but for experiment E-1

Fig. 15
figure 15

Time series of bed shear stress (Pa) at selected offshore locations (xo) for the experiments E-1 and E-2

Fig. 16
figure 16

Nonhydrostatic model simulation. Spatial distributions of the concentration field C after a 1.6 h, b 2.5 h, and c 4.3 h of simulation. The model is initialized with values from the shelf model (control experiment) at xo = 12 km after 5 days of simulation

Fig. 17
figure 17

Same as Appendix Fig. 16, but initialized with values from the shelf model (control experiment) at xo = 30 km. Shown are the distributions of C after a 1.6 h and b 6 h of simulation

Fig. 18
figure 18

Results for the nonhydrostatic model simulation that is initialized with values from the shelf model (control experiment) after 5 days of simulation. a Time series of bed shear stress (Pa) at xo = 12 km. The solid line shows the spatial average, the dashed lines account for twice the standard deviation (i.e., 95% confidence interval). Note that the upper curve roughly corresponds to maximum bed shear stresses. b Resultant average and range (based on twice the standard deviation) of bed shear stresses as a function of offshore distance xo. Arrows indicate instances in which the nonhydrostatic model predicts the onset of shear flow instabilities. The red arrow (b) highlights the region of peak bed shear stresses

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kämpf, J. Extreme bed shear stress during coastal downwelling. Ocean Dynamics 69, 581–597 (2019). https://doi.org/10.1007/s10236-019-01256-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10236-019-01256-4

Keywords

Navigation