or collective redistirbution other or collective or means this reposting, of machine, is bySociety, photocopy article only anypermitted of withSociety.Send portion theall PO B approvalOceanographycorrespondence to: [email protected] The of or Th e in Oceanography published been This has article SPECIAL ISSUE ON Coastal Ocean Processes Turbulence, Transfer, and Exchange

By James N. Moum, Jonathan D. Nash, , VolumeSociety.Society. 4, a quarterlyOceanography 21, Number The Permission journalCopyrightOceanography of All grantedrights 2008 is by The reserved. use in tocopy article teaching thisand for research. R and Jody M. Klymak Small-Scale Processes in the Coastal Ocean

20 depth (m)

40 ox 1931, R ockville, MD 20849-1931, USA. ockville, epublication, systemmatic reproduction,

-200 -100 0 100 distance along ship track (m)

22 Oceanography Vol.21, No.4 ABSTRACT. Varied observations over Oregon’s continental shelf illustrate the

beauty and complexity of geophysical flows in coastal waters. Rapid, creative, and

sometimes fortuitous sampling from ships and moorings has allowed detailed looks

at boundary layer processes, internal waves (some extremely nonlinear), and coastal

currents, including how they interact. These processes drive turbulence and mixing

in shallow coastal waters and encourage rapid biological responses, yet are poorly

understood and parameterized. The work presented here represents examples of

Figure 1. Kelvin-Helmholtz billows efforts by many physical oceanographers to quantify small-scale, coastal-mixing recently generated in a 30-m amplitude nonlinear internal wave propagating processes so that their effects might be included in regional circulation models. from left to right. The bright scattering layers trailing the wave trough are likely due to sound-speed microstructure. Such clear visualization of coherent internal structure is rare. It has also led to reasonable speculation that the deeper scattering layers in this image are associ- INTRODUCTION to a better appreciation for the role of ated with a similar phenomenon, but of Coastal oceans have a wide range small-scale processes in the coastal smaller scale and more explosive growth. Adapted from Moum et al., 2003 of fascinating and complex fluid ocean. Such efforts have allowed for motions that can only be captured by quantitative assessment of turbulence- detailed sampling. The fluid dynamics closure models for some specific cases, represented can such as the tidally mixed Celtic Sea • be beautiful to behold, as depicted (i.e., Simpson et al., 1996) and bot- in Figure 1, which shows growing tom boundary-layer turbulence on the Kelvin-Helmholtz billows within a Oregon shelf (i.e., Kurapov et al., 2005b), propagating internal solitary wave— and they have offered the potential for an example of waves within waves; improved turbulence parameterizations • excite rapid biological responses; (i.e., MacKinnon and Gregg, 2003a). • create irreversible transports of mass, Here, we provide examples from our biomass, and nutrients and momen- work over Oregon’s continental shelf. tum that are difficult to parameterize in coastal circulation models. NEW PERSPECTIVES ON CROSS-SHELF STRUCTURE Because of the multiplicity of small- It is difficult to obtain a satisfying mea- scale processes and their complex nature, sure of the detailed structure of coastal the prospects for sorting out their con- flows because tributions to the larger-scale coastal cir- • the structure is three dimensional; culation has seemed daunting. However, • the scales are O(10 km) in the cross- over the past decade, significant progress shelf dimension and larger in the has been made in understanding fine- alongshore dimension; scale and turbulence dynamics over con- • they evolve on a range of time scales, tinental shelves. Targeted efforts using including that of changes in the sur- modern, rapid-sampling shipboard and face wind field, and from both obser- bottom-mounted instrumentation have vational and dynamical perspectives, 200 provided detailed observations leading on tidal and near-inertial time scales.

Oceanography December 2008 23 To obtain detailed observations of 2001 (during the CoOP-sponsored BOTTOM cross-shelf structure, perhaps the best COAST program). Density/turbulence Although several important aspects we can do is to sample a fixed line across profiles were made directly from near of the coastal jet and the Ekman layer the shelf as rapidly as possible and the surface to the seafloor (within 2 cm appear in Figure 2, we focus first on the consider the hypothesis that the flow is of the bottom), allowing a detailed look fluid in the bottom boundary layer and two dimensional. A sequence of cross- at the structure and evolution of not interpret its motion in terms of basic shelf transects at 11–20-h intervals were only the coastal jet but also the bottom Ekman dynamics, which dictates that made with this in mind in late spring Ekman layer (Figure 2). the cross-shelf movement is determined

Figure 2. Evolution of the flow on the shelf showing coastal , its relaxation, and resumption. Summary of alongshore currents (V, lefthand column), turbulent dissipa- tion rate (ε, middle column) and turbidity (880-nm optical backscatter measurements, right column) from an eight- day period in May/June 2001 during which 12 transects were made across the Oregon shelf off Cascade Head. Isopycnals are plotted over each image, the deepest isopycnal highlighted as an indicator of cross-shelf motion of the bottom boundary layer. Relative aver- aged over the 24-h period pre- ceding each transect is shown to the right (upwelling-favorable down). The relative time of each transect is shown in the leftmost column, starting with 0 h at the beginning of the first transect. Adapted from Perlin et al., 2005a

24 Oceanography Vol.21, No.4 x V0 solely by the bottom stress in the along- shelf direction (north-south off Oregon). The densest fluid is outlined by the uEkman deepest of the three isopycnals shown in Figure 2. As the wind (and current) remains steady and southward in the first 80 h, this fluid moves up the slope at speeds of several kilometers per day (1 km d-1 ≅ 1 cm s-1). As the wind relaxes and turns to the north, the current also reverses, and the densest fluid flows back down the slope, returning up the slope with resumption of southward winds (and southward currents). Predicting the requires an estimate of the bottom stress. At this location, a typi- cal law-of-the-wall parameterization did not yield bottom stresses in agreement with either the observed Ekman trans- Figure 3. Convectively driven mixing in the bottom boundary (from data at +118 h in Figure 2). Density ports or turbulence measurements of is color-imaged in the upper panel (darker is heavier fluid). Isopcynals are shown in black and the height bottom stress (Perlin et al., 2005b). This of the bottom mixed layer in white. The upwelling-induced cross-shelf circulation has drawn dense fluid up the shelf. Upon relaxation and reversal of the winds (+98 h, +118 h), the near-bottom alongshore situation has led to a modified bottom flow V( 0) reverses, inducing a cross-shore Ekman velocity profile similar to that shown, in which the bot- stress parameterization that accounts for tom velocity must equal 0. This velocity profile combined with the induced cross-shore density gradient the suppression of turbulence by stratifi- causes light fluid to be advected offshore beneath denser fluid above. The result is convectively driven mixing, depicted in the bottom panel, which shows vertical profiles of potential temperature θ( ) and cation (Perlin et al., 2005a). ε (shading) as functions of height above bottom (hab); the ordinate label, hab, refers to height above An important wrinkle in this straight- bottom. Numerous instantaneous examples of warm (light) fluid beneath cool (heavy) fluid can be seen. forward interpretation occurs as upwell- This instability acts to drive the near-bottom turbulence and thicken the bottom Ekman layer.Adapted from Moum et al., 2004 ing relaxes (+98 h, +118 h). Individual vertical profiles indicate relatively light parcels of fluid near the seafloor at These unstable profiles are accompa- Oregon is northward (see Figure 3). As a midshelf, producing a statically unstable nied by high near-bottom turbulence, result of convectively driven near-bottom state (Figure 3; Moum et al., 2004). quantified by the viscous rate of dis- mixing, the greatest observed values of ε sipation of turbulent kinetic energy, ε. and thickest bottom boundary layers are James N. Moum (moum@coas. On average, both density and ε profiles found during these upwelling relaxation oregonstate.edu) is Professor, College of closely resemble those observed during events, providing a clear explanation for Oceanic and Atmospheric Sciences, Oregon convectively driven mixing in the upper the thickened boundary layers previously State University, Corvallis, OR, USA. ocean in response to surface cooling or observed elsewhere under similar condi- Jonathan D. Nash is Associate Professor, in the lower atmosphere when heated tions (Trowbridge and Lentz, 1991). This College of Oceanic and Atmospheric from below. The manner by which this process has been implicated in oceanic Sciences, Oregon State University, occurs in the bottom Ekman layer is frontogenesis (Thomas and Lee, 2005) Corvallis, OR, USA. Jody M. Klymak is thought to be a result of the combination and in the formation of slippery bot- Assistant Professor, School of Earth and of the cross-shelf density gradient and tom boundary layers predicted from Ocean Sciences, University of Victoria, the shape of the Ekman velocity profile laboratory and numerical experiments Victoria, BC, Canada. when the flow atop the bottom layer off (MacCready and Rhines, 1993).

Oceanography December 2008 25 NONLINEAR INTERNAL WAVES 2002; Klymak and Moum, 2003). The imaging tool (Farmer and Smith, 1980). One clearly visible small-scale process leading edges of these waves represent The passage of the waves causes defor- in coastal oceans is revealed through convergence zones, where the ocean mation of preexisting scattering layers, long (tens of kilometers) streaks on surface is ruffled (providing the visible and under the right conditions, the wave the sea surface. These streaks are often signature) and fluid is displaced down- itself can create scattering layers through associated with nonlinear internal waves, ward (providing a supply of near-surface internal instabilities that create small- sometimes referred to as solitary waves material to subsurface predators). The scale density microstructure and sound- (even though they are rarely solitary), local food chain can be extensive and speed fluctuations. Figure 1 is a dramatic or solitons. They are visible from air active, marked by tuna swarms and dol- example of this phenomenon. In this and space, although it is the compila- phins hunting along the wave fronts, a image, numerous individual scatterers tions from satellite synthetic aperture phenomenon long known by local fisher- are evident, likely small fish or zooplank- radar observations that provide the best men. Less well known are the biological ton, as well as coherent scattering layers evidence for their constant presence in responses that are internal and rapid, that describe the outline of the wave and coastal oceans (http://internalwaveatlas. as can be seen in space/time series of that are thought to be created by small- com/Atlas2_index.html). Although these acoustics (Figure 4). scale internal deformation within the streaks are referred to as waves, part of Only recently have measurements of wave (as in the upper panels of Figure 5; their highly nonlinear nature may make the internal structure of the waves been Moum et al., 2003). them more like vortices, in the sense that of sufficient fidelity to quantify funda- Shallow pycnoclines are present in fluid may be trapped internally as the mental aspects of their dynamics. One spring and summer off the Oregon coast waves propagate, thereby transporting means of measuring the wave structure (Figure 5). They act as waveguides along fluid across the continental shelf (Lamb, uses high-frequency acoustics as an which waves of depression propagate,

Acoustic Backscatter 0

20

40 ]

60 DEPTH [m 80

100

120 13:55 14:02 14:09 14:16 14:24 14:31 14:38 TIME AUG 21, 2001 Figure 4. Biological reaction to passing solitary waves on the Oregon Shelf. The image was constructed from hull-mounted acoustic backscatter measurements (120 kHz) over Hecate Bank. The seafloor is seen as a hard (red) target at 112-m depth. Near-surface scat- tering (likely from zooplankton) shows the passage of a sequence of nonlinear internal waves of depression that propagate along a near-surface thermocline, the first at about 14:12. Prior to the passage of the first wave, there is a relatively horizontal scattering hori- zon at 60-m depth. Following this, the layer appears to oscillate vertically, moving upward to meet the second wave at 14:20. If we can consider this layer as a contiguous assemblage of zooplankton or small fish, they appear to have swum upward across isopycnals (not shown here) very rapidly (tens of cm s-1) through the phase of the wave with downward vertical velocities. This example is typical of many showing upward motion of deep scattering layers apparently triggered by the passage of nonlinear internal waves.

26 Oceanography Vol.21, No.4 Figure 5. Seasonal differences in nonlinear internal waves. Summertime stratification over Oregon’s continental shelf provides a near-surface waveguide that supports waves of depression, examples of which are shown in the uppermost panel. Many of the waves’ properties can be measured from ship. High-frequency acoustics helps visualize the waves’ structure. The dots represent times at which vertical profiles of density and turbulence were obtained through an isopyc- nal. The temperature structure immediately above and below that isopycnal is indicated in the second panel, beneath which are shown the temperature fluctuations and high values of the turbulence kinetic energy dissipation rate (ε) due to the passage of the waves. In contrast, wintertime stratification provides a near-bottom waveguide that sup- ports waves of elevation. Although these waves can also be measured from ship, they are perhaps best observed from sensors mounted on the seafloor. The lower panels show records from an upward-looking acoustic Doppler current profiler and an acoustic Doppler velocimeter, both mounted at 1-m height above the bottom. Horizontal velocity is in the direction of wave propagation (roughly onshore). The veloc- ity at 1-m height is derived from the velocimeter measure- ments and the turbulence dissipation at 1-m height by scaling velocity spectra in the inertial subrange. Adapted from Moum et al., 2003; Moum et al. 2007

Oceanography December 2008 27 their surface signatures made visible by of depression. Instead, near-bottom is comparatively weak and rarely seen the patterns of alternating, near-surface thermoclines permit the propagation (Zhao et al., 2003). These waves are best fluid convergence/divergence. In winter, of near-bottom waves of elevation. sampled from below, where stationary, near-surface stratification is typically Although these waves might exhibit a bottom-mounted landers with rapidly weak and unable to support waves surface signature in shallow water, it sampled acoustic Doppler current

VARICOSE WAVES By Emily Shroyer

Nonlinear internal waves of depression and elevation are often seen in the coastal ocean. The terms depression/elevation imply waves that have a mode-1 structure (i.e., all isopycnals are displaced in the same direction, either down or up). Higher-mode waves also exist. Here, we see an example of a wave with mode-2 structure (A) propagating onshore (to the right) followed by a train of short, mode-1 depression waves (B). In the mode-2 wave, isopycnals near the surface are displaced upwards while those at depth are displaced downwards. These waves are sometimes referred to as “varicose” to emphasize the “bulging” nature of the isopycnals. In the middle panel, the back- ground velocity has been removed, leaving the veloc- ity field attributed to the wave. Note the onshore (red) pulse centered at approximately 12 m and the offshore (blue) return flow near the surface. Any return flow beneath the onshore pulse is weak and undetectable in our measurements. Although varicose waves are not as energetic as larger depression waves observed in the same region, they potentially play a significant role in mixing shelf waters. These observations show elevated turbulent kinetic energy dissipation rates (approaching 10-4 m2 s-3) at the trailing edge of the mode-2 wave. Considering the location of these waves in the middle of the pycnocline, this intense mixing may play an important role in eroding the barrier established by the pycnocline, thereby enhancing ver- tical fluxes of nutrients and heat in the water column.

Figure B1. Upper panel: Acoustic backscatter. Middle panel: Emily Shroyer ([email protected]) is Onshore wave velocity, U, with isopycnals contoured. Lower PhD Candidate, College of Oceanic and Atmospheric panel: Turbulence dissipation rate, from shipboard measure- ments over the New Jersey shelf in 85-m water depth. Shoreward Sciences, Oregon State University, Corvallis, OR, USA. direction is to the right.

28 Oceanography Vol.21, No.4 profilers (ADCPs) can provide continu- nonlinear wave fields on continental mountain range), circulation may be ous, high-resolution measurements over shelves. Although this would seem to be blocked—the flow is diverted either over long time periods. An example is seen in an easily tested mechanism, wave arriv- or around the obstacle. Consequences in the lower few panels of Figure 5, where als at a fixed observation point are rarely coastal flows include accelerated near- the leading three waves have amplitudes regular, probably because variations in bottom velocities (causing enhanced of 30 m. Within each of these waves, mesoscale stratification and currents on bottom stress) and a pressure drop across near-bottom ε is considerably larger the shelf alter not only generation loca- the obstacle, resulting in form drag. The than ambient—either turbulence is tions and timing, but wave propagation exact response is determined by the explosively generated within the waves as well. The finding that these waves can topographic geometry, local stratifica- or it is being advected along with the be simply detected by seafloor pressure tion, and flow speed. An unambiguous wave. The distinction is a matter of some measurement (Moum and Nash, 2008) example of an observed blocked flow importance. If turbulence is advected by provides some hope for inexpensive, over a small bank on Oregon’s continen- the waves, then fluid must be trapped long-term monitoring to assess wave tal shelf shows some of the main features in the waves and advected with them prevalence over continental shelves. (Figure 6; Nash and Moum, 2001). The like a turbulent bore on a beach. This drag exerted by this 2-km feature exerts trapped fluid would include the resus- FORM DRAG the equivalent of O(50 km) of flat- pended sediment load created by the One aspect of coastal flows that distin- bottom stress. Such clear measurement of large bottom stresses as the waves pass guishes them from those in the main these flows requires intensive shipboard over the bottom (which can be measured oceanic gyres is their proximity to the profiling efforts and, at present, there optically by profilers—as in Figure 2). bottom. In this sense, coastal flows feel exist only anecdotes. To determine the Evidence for trapped cores has come the bottom in a manner similar to the net effect of these flows on coastal cir- from observations that may indicate way that atmospheric circulation does: culation requires longer-term measure- zooplankton or small fish clouds wholly when atmospheric circulation encounters ments combined with modeling. One contained within the leading wave of a a particularly large obstacle (such as a way to do this is by directly measuring near-bottom wave train propagating up the shelf (Klymak and Moum, 2003). Any onshore advection of fluid by eleva- tion waves during periods of coastal downwelling counters the prevailing Figure 6. High-drag flow over a bottom Ekman transport; it is possible topographic obstruction in the coastal ocean. The figure shows the structure that this is an efficient means of intermit- of density and ε over Stonewall Bank, a tently injecting near-bottom fluid from small rise on Oregon’s continental shelf. offshore onto the shelf. The asymmetry in density structure indicates flow controlled by internal An outstanding question is how hydraulics, here characterized by a nonlinear internal waves are formed. near-bottom flow that is accelerated One mechanism, reported by Nash and across the bank from right to left, caus- ing high ε in both the bottom bound- Moum (2005), shows wave fission from ary layer and the sheared layer above. the decelerating, buoyant Columbia The near-bottom flow terminates downstream in a turbulent hydraulic River plume. However, this process is jump. This flow exerts high drag on the local to the plume, and yet these waves overlying flow in the form of intense are ubiquitous on the shelf. We suspect bottom stresses (skin friction) as well as a net pressure drop (form drag) that nonlinear steepening of internal across the bank. Adapted from Nash tides generated at shelf breaks (Holloway and Moum, 2001 et al., 1997) is primarily responsible for

Oceanography December 2008 29 the pressure drop across the bank with layers in the mesosphere, where they wave field is also not typically large high-resolution seafloor pressure sensors, deposit their momentum. However, the enough to trigger shear instability on an effort now underway. free surface of the ocean acts as a reflec- its own. However, enhanced turbulence As a consequence of form drag, tor, and we do not yet understand how is observed where the superposition of momentum is extracted from the mean momentum is removed via wave radia- thermal wind shear and internal gravity flow—vfrom both the pressure drop tion from features like Stonewall Bank or wave shear reduces gradient Richardson across the obstacle and the enhanced where it is deposited. numbers to values that are favorable bottom stress. Upstream flow blocking to instability (Figure 7; Avicola et al., and the release of downstream lee waves THE ROLE OF INTERNAL WAVES 2007). Because the coastal jet is in geo- during unsteady flows can produce IN SHEAR-ENHANCEMENT OF strophic balance to first order, inclusion nonlocal consequences and stresses, THE COASTAL JET of a coastal internal wave field is not as the resultant radiated waves can Away from surface and bottom bound- necessary for a first-order accounting transport momentum elsewhere. In the ary layers, interior turbulence is strongly of the coastal circulation. However, lower atmosphere, where a consider- linked to the structure of the upwelling- the lack of shear in the interior of the able fraction of the momentum is lost induced coastal jet, and especially to geostrophic flow means that it cannot to mountain drag, an accounting of this locations where the jet shear is large reasonably predict large mixing (or wave radiation through parameterization (Figure 2). Yet, the thermal wind shear large ε). The internal wave field must is an important part of global circula- alone is of insufficient intensity to trig- also be captured, or at least parameter- tion models. It is not yet clear how best ger shear instability. Continental shelves ized. It is possible that jet-wave interac- to reconcile the analogous processes in have a broad spectrum of internal tions are not simply linear, so a simple coastal flows, because we do not know waves, with peaks at near-inertial and superposition of constituent shear is a the partitioning of local and remote tidal periods (Levine, 2002; MacKinnon naïve approach to assessing potential momentum sinks. Atmospheric waves and Gregg, 2003b; Avicola et al., 2007); for instability. The prospects for non- propagate vertically and break at critical the shear associated with this internal linear interactions between the coastal

Figure 7. Thermal wind Richardson number (Ri) compared to observed Ri and ε, each computed from data corresponding to the cross-shelf transect marked +197 h in Figure 2. The left panel shows inverseRi computed using thermal wind shear alone, the middle panel inverse Ri computed from the observed shear (4-m scale) and the right panel the observed ε. The critical component of the observed shear is that due to internal waves. Clearly, the thermal wind shear does not display the structure that is observed in the turbulence, ε. Adapted from Avicola et al., 2007

30 Oceanography Vol.21, No.4 Turbulence in River Plumes By Levi Kilcher

Shear, stratification, and turbulence associated with tidally driven river plumes far exceed that associated with typical coastal flows, except perhaps during major storms. Acoustic images along an east-west transect across the Columbia River plume reveal the complex structure typical of each tidally pulsed plume. On the ebb of each tidal cycle, a pulse of low- salinity water is released from the river mouth, which in this case detaches from the bottom about 8 km from the river’s mouth. Out to about 11.5 km, the near-surface flow thins from 20- to 10-m thickness, is continuously stratified, and has rela- tively uniform shear and acoustic backscatter. At that point, it transitions: shear, stratification, and turbulence become concentrated at the base of the surface layer (visible in both acoustic backscatter and S2). Farther offshore, the plume continues to thin, culminating in a 20-m-deep depression at the plume front, which propagates westward at approximately 1 m s-1. Energy dissipation rates within the turbulent head can exceed 10-3 m2 s-3, considerably greater than that within dissipating nonlinear internal waves in the coastal ocean (Moum et al., 2003). The plume front is also a region of strong Figure B2. Acoustic backscatter (upper panel), shear squared (middle), and along-axis velocity (lower) within horizontal velocity convergence, trapping buoyant particles, the tidally pulsed plume of the Columbia River in August steepening local surface waves, and potentially enhancing 2005. Grey bars in the bottom panel show turbulence dis- biological productivity. As the plume front decelerates, it sipation rate (scale below). Distances are referenced to the river’s mouth to the east. can release large-amplitude internal waves (Nash and Moum, 2005), which may propagate 50 km offshore, further extending the influence of the river plume.

Levi Kilcher ([email protected]) is PhD Candidate, College of Oceanic and Atmospheric Sciences, Oregon State University, Corvallis, OR, USA.

jet and internal gravity waves have shear-driven mixing associated with the CONSEQUENCES yet to be investigated. superposition of coastal jet and internal Detailed observations of coastal pro- Significantly, because the vertical waves, these transports will be underesti- cesses reveal an impressive variety of redistribution of scalars such as heat, mated if the turbulence parameterization phenomena. Although it is clear that biomass, and nutrients is caused by is based on the subtidal flow alone. these phenomena are all at least locally

Oceanography December 2008 31 important, their role in shaping the char- may be indicated by comparing the tur- and associated turbulence may fail to acter of the low-frequency flow remains bulence dissipation rate (ε) as observed reproduce the correct structure and/or an open question. In fact, these observa- at +0 h in Figure 2 to that calculated transport of the coastal jet. tions are, to a large degree, anecdotal. through a turbulence model embedded The possible importance of form drag Without a better idea of their prevalence, in the Princeton Ocean Model (POM), is suggested by a different numerical it will be impossible to determine their used to compute turbulence fluxes over experiment. The correction term in data- importance. Extended, targeted obser- the same time period (Figure 8; Kurapov assimilative models applied off Oregon vations will be needed to assess their et al., 2005b). The model clearly devel- opposes the direction of the mean flow, ubiquity and the prospects for incorpo- ops turbulence in surface and bottom in a depth-averaged sense (Figure 9; rating their dynamics into large-scale boundary layers, although the structure Kurapov et al., 2005a). In this case, the models. But, with some knowledge of (i.e., boundary layer location and thick- required correction is largest in regions their properties and of recent numerical ness) differs from the observations. But of shallow and rough topography, and experiments, maybe we can identify a perhaps more importantly, the interior is of the same order of magnitude as couple of significant departures of model is devoid of turbulence, possibly because the computed bottom stress. While it from observation. the model does not contain an energetic is often argued that the correction may Low-resolution, low-frequency mod- internal wave field (high-resolution, two- be associated with a missing large-scale els do not resolve form drag, nor do they dimensional model studies by Federiuk pressure gradient, the spatial structure develop sufficient shear to generate the and Allen [1996] have produced interior of the correction appears linked to the turbulence observed in the interior of turbulence caused by inertial wave topography. If the missing physics here coastal upwelling systems. And non- shear via turbulence parameterization). is form drag, and if it can be properly assimilative coastal models can overesti- Because the high observed turbulence parameterized numerically, perhaps a mate flow speeds. Could these symptoms levels at 30- to 50-m depth inshore of significant part of the correction term result from neglect of the effects of 10 km are in fact coincident with large can be eliminated and the accuracy small-scale processes? velocity gradients at the base of the jet, of forward models increased without The role of the internal wave field models that lack an internal wave field data assimilation. Parameterizing and incorporating these processes into the next generation of coastal models ought to be of high priority but will require both significant ingenuity and further extended, detailed measurements as well as enhanced collaborations between observationalists and modelers.

ACKNOWLEDGEMENTS The work described here was funded by the Office of Naval Research and the National Science Foundation (includ- ing CoOP projects), and the authors continue to benefit from the funding of these organizations. The efforts and lead- ership of many fellow scientists, includ- ing postdoctoral fellows and students Figure 8. Comparison of observed (a) and modeled (c) turbulence dissipation rate (ε) across the Oregon shelf corresponding to the +0 h transect in Figure 2. The model-derivedε incorporates data assimilation whose names appear in the list of refer- of mean flow variables, but the model does not include the tides.Extracted from Kurapov et al., 2005b ences, have been critical to our advances

32 Oceanography Vol.21, No.4 MacKinnon, J.A., and M.C. Gregg. 2003a. Mixing on the late-summer New England shelf—Solibores, shear, and stratification. Journal of Physical Oceanography 33:1,476–1,492. MacKinnon, J.A., and M.C. Gregg. 2003b. Shear and baroclinic energy flux on the summer New England shelf. Journal of Physical Oceanography 33(7):1,462–1,475. Moum, J.N., and J.D. Nash. 2008. Seafloor pres- sure measurements of nonlinear internal waves. Journal of Physical Oceanography 38(2):481–491, doi:10.1175/2007JPO3736.1. Moum, J.N., J.M. Klymak, J.D. Nash, A. Perlin, and W.D. Smyth. 2007a. Energy transport by nonlinear internal waves. Journal of Physical Oceanography 37(7):1,968–1,988. Moum, J.N., A. Perlin, J.M. Klymak, M.D. Levine, T. Boyd, and P.M. Kosro. 2004. Convectively driven mixing in the bottom boundary layer. Journal of Physical Oceanography 34:2,189–2,202. Moum, J.N., D.M. Farmer, W.D. Smyth, L. Armi, and S. Vagle. 2003. Structure and generation of turbulence Figure 9. The alongshore momentum correction term required in data-assimilation at interfaces strained by internal solitary waves models is of the same sign and magnitude as the bottom stress, suggesting that an propagating shoreward over the continental shelf. additional momentum sink is not incorporated. This figure shows a comparison Journal of Physical Oceanography 33:2,093–2,112. of modeled bottom stress (left) and the required correction term (right) in the Nash, J.D., and J.N. Moum. 2001. Internal hydraulic Oregon shelf depth-averaged momentum budget; black lines are isobaths. Because flows on the continental shelf: High drag states the correction is highest in regions of shallow water and rough topography, we sug- over a small bank. Journal of Geophysical Research gest that missing small-scale physics (in particular, form drag) may be responsible 106:4,593–4,611. for the required correction. Extracted from Kurapov et al., 2005a Nash, J.D., and J.N. Moum. 2005. River plumes as a source of large-amplitude internal waves in the coastal ocean. Nature 437:400–403, doi:10.1038/ nature03936 Perlin, A., J. N. Moum, and J.M. Klymak. 2005a. in understanding the physical nature transformation on the Australian North West Shelf. Response of the bottom boundary layer over a Journal of Physical Oceanography 27:871–896. sloping shelf to variations in alongshore wind. of these small-scale flows and how they Klymak, J.M., and J.N. Moum. 2003. Internal soli- Journal of Geophysical Research 110(C10S09), may contribute to the larger coastal cir- tary waves of elevation advancing on a sloping doi:10.1029/2004JC002500. culation. We are, as always, indebted to shelf. Geophysical Research Letters 30(20):2,045, Perlin, A., J.N. Moum, J.M. Klymak, M.D. Levine, T. doi:10.1029/2003GL017706. Boyd, and P.M. Kosro. 2005b. A modified law-of- the outstanding technical support staff of Kurapov, A.L., J.S. Allen, G.D. Egbert, R.N. Miller, the-wall to describe velocity profiles in the bottom the Ocean Mixing Group at Oregon State P.M. Kosro, M.D. Levine, and T. Boyd. 2005a. boundary layer. Journal of Geophysical Research Distant effect of assimilation of moored currents 110(C10S10), doi:10.1029/2004JC002310. University. Jane Huyer and John Allen into a model of coastal wind-driven circula- Simpson, J.H., W.R. Crawford, T.P. Rippeth, A.R. contributed helpful comments. tion off Oregon. Journal of Geophysical Research Campbell, and J.V.S. Choek. 1996. The vertical 110(C02022), doi:10.1029/2003JC002195. structure of turbulence dissipation in shelf seas. Kurapov, A.L., J.S. Allen, G.D. Egbert, R.N. Miller, P.M. Journal of Physical Oceanography 26:1,579–1,590. References Kosro, M.D. Levine, T. Boyd, and J.A. Barth. 2005b. Thomas, L.N., and C.M. Lee. 2005. Intensification Avicola, G., J.N. Moum, A. Perlin, and M.D. Levine. Assimilation of moored velocity data in a model of ocean fronts by down-front winds. Journal of 2007. Enhanced turbulence due to the superposi- of coastal wind-driven circulation off Oregon: Physical Oceanography 35:1,086–1,102. tion of internal gravity wave shear on the coastal Multivariate capabilities. Journal of Geophysical Trowbridge, J.H., and S.J. Lentz. 1991. Asymmetric upwelling jet. Journal of Geophysical Research Research 110(C10S08), doi:10.1029/2004JC002493 behavior of an oceanic boundary layer over a 112(C06024), doi:10.1029/2006JC003831. Lamb, K.G. 2002. Shoaling nonlinear internal sloping bottom. Journal of Physical Oceanography Farmer, D.M., and J.D. Smith. 1980. Tidal interaction waves: On a criterion for the formation of waves 21:1,171–1,185. of stratified flow with a sill in Knight Inlet. Deep- with trapped cores. Journal of Fluid Mechanics Zhao, Z., V. Klemas, Q. Zheng, and X. Yan. 2003. Sea Research 27:239–245. 478:81–100. Satellite observations of internal solitary waves Federiuk, J., and J.S. Allen. 1996. Model studies of Levine, M.D. 2002. A modification of the Garrett– converting polarity. Geophysical Research Letters near-inertial waves in flow over the Oregon con- Munk internal wave spectrum. Journal of Physical 30, doi:10.1029/2003GL108286. tinental shelf. Journal of Physical Oceanography Oceanography 32:3,166–3,181. 26:2,053–2,075. MacCready, P., and P.B. Rhines. 1993. Slippery bottom Holloway, P.E., E. Pelinovsky, T. Talipova, and B. boundary layers on a slope. Journal of Physical Barnes. 1997. A nonlinear model of internal tide Oceanography 23:5–22.

Oceanography December 2008 33