Surface waves and spatially coherent structures in the near-surface layer of the ocean

Similar documents
The Near-Surface Layer of the Ocean

Analysis of Near-Surface Oceanic Measurements Obtained During CBLAST-Low

Roughness Sub Layers John Finnigan, Roger Shaw, Ned Patton, Ian Harman

Northern Arabian Sea Circulation Autonomous Research (NASCar) DRI: A Study of Vertical Mixing Processes in the Northern Arabian Sea

Parameterizing the Effects of Upper-Ocean Large Eddies on Air-Sea Interaction

Ocean Surface Mixed Layer

Ocean Mixing and Climate Change

Impact of Typhoons on the Western Pacific Ocean (ITOP) DRI: Numerical Modeling of Ocean Mixed Layer Turbulence and Entrainment at High Winds

Eddy viscosity. AdOc 4060/5060 Spring 2013 Chris Jenkins. Turbulence (video 1hr):

2013 Annual Report for Project on Isopycnal Transport and Mixing of Tracers by Submesoscale Flows Formed at Wind-Driven Ocean Fronts

Wind driven mixing below the oceanic mixed layer

Characteristics of Langmuir Turbulence in the Ocean Mixed Layer

Impacts of Ocean Waves on the Atmospheric Surface Layer: Simulations and Observations

Large-Eddy Simulations of Tropical Convective Systems, the Boundary Layer, and Upper Ocean Coupling

Observations of wind, waves, and the mixed layer: the scaling of surface motion*

Scaling Kinetic Energy in the Wind-Driven Mixed Layer

Synthesis of TKE Dissipation Rate Observations in the Ocean's Wave Zone

Temperature fronts and vortical structures in turbulent stably stratified atmospheric boundary layers

Hurricanes are intense vortical (rotational) storms that develop over the tropical oceans in regions of very warm surface water.

meters, we can re-arrange this expression to give

An Introduction to Coupled Models of the Atmosphere Ocean System

Effects of transfer processes on marine atmospheric boundary layer or Effects of boundary layer processes on air-sea exchange

Prototype Instabilities

Resonant Wind-Driven Mixing in the Ocean Boundary Layer

Understanding Near-Surface and In-Cloud Turbulent Fluxes in the Coastal Stratocumulus-Topped Boundary Layers

Upper Ocean Circulation

Surface waves and oceanic mixing: Insights from numerical simulations with stochastic surface forcing

Donald Slinn, Murray D. Levine

Wind-driven Western Boundary Ocean Currents in Terran and Superterran Exoplanets

Large-Eddy Simulations of Tropical Convective Systems, the Boundary Layer, and Upper Ocean Coupling

SMS 303: Integrative Marine

The Planetary Circulation System

Upper-Ocean Processes and Air-Sea Interaction in the Indonesian Seas

REPORT DOCUMENTATION PAGE

Role of Turbulent Mixing in Plankton Dynamics Simulated by Large Eddy Simulation (LES) YIGN NOH Yonsei University, S. Korea

Lecture 3. Turbulent fluxes and TKE budgets (Garratt, Ch 2)

PUBLICATIONS. Journal of Geophysical Research: Oceans

Logarithmic velocity profile in the atmospheric (rough wall) boundary layer

Analysis of Mixing and Dynamics Associated with the Dissolution of Hurricane-Induced Cold Wakes

Numerical Simulation of the Wind Stress Effect on ALOS PALSAR Images of Far Wakes of Ships Atsushi Fujimura and Alexander Soloviev

The Stable Boundary layer

LECTURE 28. The Planetary Boundary Layer

6A.3 Stably stratified boundary layer simulations with a non-local closure model

Double-diffusive lock-exchange gravity currents

This is the first of several lectures on flux measurements. We will start with the simplest and earliest method, flux gradient or K theory techniques

Internal Wave Driven Mixing and Transport in the Coastal Ocean

A note on mixing due to surface wave breaking

Turbulent boundary layer

The influence of crosswind tidal currents on Langmuir circulation in a shallow ocean

Fluid Mechanics. Chapter 9 Surface Resistance. Dr. Amer Khalil Ababneh

BALANCED FLOW: EXAMPLES (PHH lecture 3) Potential Vorticity in the real atmosphere. Potential temperature θ. Rossby Ertel potential vorticity

Air-sea Gas Exchange and Bio-surfactants: Low and High Wind Speed Extremes

ESCI 110: 2 s.h. Introduction to Earth Sciences Programs ESCI 322: 3 s.h. Environmental Hydrology ESCI 241: 4 s.h. Meteorology (G2, L)

Contents. Parti Fundamentals. 1. Introduction. 2. The Coriolis Force. Preface Preface of the First Edition

Boundary layer processes. Bjorn Stevens Max Planck Institute for Meteorology, Hamburg

The Effect of Rainfall on the Surface Layer during a Westerly Wind Burst in the Western Equatorial Pacific

Contribution of coherent vortices such as Langmuir cells to wind-driven surface-layer mixing

Lecture 9: Tidal Rectification, Stratification and Mixing

LES of turbulent shear flow and pressure driven flow on shallow continental shelves.

A Numerical Study of a TOGA-COARE Squall-Line Using a Coupled Mesoscale Atmosphere-Ocean Model

Quasi-3D Nearshore Circulation Equations: a CL-Vortex Force Formulation

Baroclinic Frontal Instabilities and Turbulent Mixing in the Surface Boundary Layer, Part II: Forced Simulations

The Urban Canopy and the Plant Canopy

Lecture 8. Lecture 1. Wind-driven gyres. Ekman transport and Ekman pumping in a typical ocean basin. VEk

7C.6 MOMENTUM FLUX STRUCTURES AND STATISTICS IN LOW-WIND MARINE SURFACE LAYERS: OBSERVATIONS AND LARGE-EDDY SIMULATIONS

3-Fold Decomposition EFB Closure for Convective Turbulence and Organized Structures

A Pair of Langmuir Cells in a Laboratory Tank (I) Wind-Only Experiment

Numerical Simulation of the Response of the Ocean Surface Layer to Precipitation

Turbulence is a ubiquitous phenomenon in environmental fluid mechanics that dramatically affects flow structure and mixing.

2. Baroclinic Instability and Midlatitude Dynamics

Upper-Ocean Turbulence during a Westerly Wind Burst: A Comparison of Large-Eddy Simulation Results and Microstructure Measurements

On the Coupling of Wind Stress and Sea Surface Temperature

Lecture 14. Marine and cloud-topped boundary layers Marine Boundary Layers (Garratt 6.3) Marine boundary layers typically differ from BLs over land

Observations of Giant Bursts Associated with Microscale Breaking Waves

Ocean Dynamics. The Great Wave off Kanagawa Hokusai

Sergej S. Zilitinkevich 1,2,3. Helsinki 27 May 1 June Division of Atmospheric Sciences, University of Helsinki, Finland 2

Submesoscale Routes to Lateral Mixing in the Ocean

Passive Scalars in Stratified Turbulence

HORSEHOE VORTICES IN UNIFORMLY SHEARED TURBULENCE

18A.2 PREDICTION OF ATLANTIC TROPICAL CYCLONES WITH THE ADVANCED HURRICANE WRF (AHW) MODEL

7. Basics of Turbulent Flow Figure 1.

Turbulence - Theory and Modelling GROUP-STUDIES:

Vorticity-based Analytical Models for Internal Bores and Gravity Currents

Ocean surface circulation

Turbulence and Transport The Secrets of Magnetic Confinement

Dynamics and Kinematics

ESCI 344 Tropical Meteorology Lesson 11 Tropical Cyclones: Formation, Maintenance, and Intensification

Improving Surface Flux Parameterizations in the NRL Coupled Ocean/Atmosphere Mesoscale Prediction System

General Comment on Lab Reports: v. good + corresponds to a lab report that: has structure (Intro., Method, Results, Discussion, an Abstract would be

Geophysics Fluid Dynamics (ESS228)

TURBULENT KINETIC ENERGY

Super-Parameterization of Boundary Layer Roll Vortices in Tropical Cyclone Models

High Wind Upper Ocean Mixing with Explicit Surface Wave Processes

Subsurface Expressions of Sea Surface Temperature Variability under Low Winds

Tutorial School on Fluid Dynamics: Aspects of Turbulence Session I: Refresher Material Instructor: James Wallace

Tropical Cyclone Intensification

Footprints: outline Üllar Rannik University of Helsinki

Vortices in the ocean. Lecture 4 : Baroclinic vortex processes

The atmospheric boundary layer: Where the atmosphere meets the surface. The atmospheric boundary layer:

ADAPTATION OF THE REYNOLDS STRESS TURBULENCE MODEL FOR ATMOSPHERIC SIMULATIONS

Transcription:

Surface waves and spatially coherent structures in the near-surface layer of the ocean Alexander Soloviev 1,2 and Cayla Dean 1 1 Oceanographic Center, Nova Southeastern University, USA 2 Rosenstiel School of Marine and Atmospheric Science, University of Miami, USA 3 Department of Oceanography, University of Hawaii at Manoa, USA Introduction Spatially coherent organized structures have been recognized as an important part of turbulent boundary layer processes. Langmuir cells and ramp-like structures are believed to transfer an appreciable portion of the momentum, heat, pollutants (e.g., oil), and other substances in the upper layer of the ocean. Presence of the free surface significantly complicates the analysis of turbulent exchanges at the air-sea interface. The coherent structures are not yet completely understood. A schematic diagram of Langmuir circulation is shown in Figure 1. Langmuir cells are parallel and oriented nearly downwind with alternating longitudinal vorticity, producing convergence and divergence zones. The convergence zones are substantially narrower than the divergence zones. The spacing between the Langmuir vortices ranges from a few meters to a few hundred meters (Pollard, 1977). Figure 1. Pollard s (1977) sketch of Langmuir circulations.

Langmuir circulation can be seen on the sea surface due to collection of flotsam or foam from breaking waves in convergence zones. In Figure 2, Langmuir cells are visualized by an oil spill (Fig. 2a) and foam from breaking waves (Fig. 2b). For developed seas, direction of wind, wave, and Langmuir cells is practically the same. In general, however, they do not necessarily coincide. In the two cases shown in Figure 2, the direction of Langmuir cells is different from the wave direction. (a) (b) Figure 2. Direction of Langmuir cells does not always coincide with wave direction: (a) http://www.nwfdailynews.com; (b) http://www.esoxhunter.com/walleyewind.php#eighth The atmospheric boundary layer above the ground exhibits spatially coherent organized motions in the form of ramps (Antonia et al. 1979; Phong-Anant et al. 1980). Ramp-like structures are believed to be a wide spread feature in the upper ocean as well and have been found in the upper layer of the ocean under both stable (Thorpe and Hall 1987) and unstable (Soloviev 1990) stratification. Figure 3 illustrates schematics of ramp-like structures during unstable stratification in the upper layer of the ocean. Figure 3. Schematic diagram showing ramp-like structure during nighttime conditions (After Soloviev, 1990).

Wijesekera et al (1999) collected large statistics on ramp-like structures during TOGA COARE (Fig. 4a). These data suggested that the direction of frontal interfaces produced by ramp-like structures is approximately perpendicular to wind direction (Fig. 4b). According to the field study performed by Thorpe et al. (2003) using an autonomous underwater vehicle, both types of coherent structures could coexist. Vortices associated with ramp-like structures have transverse axes, while Langmuir circulations have longitudinal axes, relative to the wind direction. How can two organized motions with perpendicular axis coexist in 3D space? (a) (b) Figure 4. A horizontal temperature profile observed at 2 m during the night of December 31, 1992 from the R/V Wecoma moving downwind at 4 m s 1. Wind stress (westerly) is about 0.1 N m 2, net surface cooling about 250 W m 2, the stability parameter, z / 0. 1, where L O is the s L O Oboukhov length scale, and average bow sensor depth z 2 m. (After Wijesekera et al. 1999.) s Wave breaking is a powerful mechanism producing significant energy flux to small scale turbulence. Benilov and Ly (2002) conditionally divided the upper ocean turbulent boundary layer into the following three sublayers (Fig. 5): 1. The wave-stirred layer (a layer of intense mixing by breaking waves). The turbulent kinetic energy (TKE) production by wave breaking substantially exceeds the mean shear production. 2. The turbulent diffusion layer. The turbulent diffusion of TKE from the wave stirred layer exceeds the wave (as well as the mean shear) effect. 3. The wall layer. The mean shear production of turbulent energy dominates over wave breaking turbulence. A logarithmic layer can form but stratification and rotation can also be important.

Figure 5. Structure of the upper ocean turbulent boundary layer below breaking surface waves (Soloviev and Lukas, 2014). Figure 6 summarizes the results of field and theoretical studies of wave-breaking turbulence. In Figure 6, only the data obtained during high wind speeds are presented in order to minimize the influence of thermohaline stratification effects on near-surface turbulence characteristics. Most of the turbulent energy dissipation takes place within one significant wave height, while the turbulent diffusion layer extends to approximately ten significant wave heights. Figure 6. Turbulence dissipation in the upper ocean. Normalized dissipation rate εh s /F versus dimensionless depth z /H s according to field and theoretical results. Here ε is the dissipation rate of the TKE, F the flux of the kinetic energy from wind to waves, and H s is the significant wave height. Wind speed range is from 7 to 19 m s 1. (After Soloviev and Lukas, 2014).

Modeling Langmuir circulation and ramp-like structures Large eddy simulation (LES) models of Langmuir circulation are now very sophisticated but still pose some questions (Soloviev and Lukas, 2014). From the basic principles of nonlinear dissipative systems, the process of self-organization reduces chaos and dissipation in the system and increases the effectiveness of property transport. In contrast, existing LES models of Langmuir circulation demonstrate significant increase of turbulent dissipation. Also, the traditional models of Langmuir circulation do not account for ramp-like structures, which are widespread features in the upper ocean turbulent boundary layer. In traditional models, Langmuir circulation is driven by the vortex force due to Stokes drift (Craik and Lebovich, 1976): F ~ u du dz (1) v s / where u s is the Stokes drift and u is the horizontal velocity component. Due to wave stirring, the near surface layer becomes almost homogeneous, like a slab layer (Fig. 7). As a result, du / dz 0. (2) Thus, the vortex force F v may vanish in that layer due to wave-breaking mixing. Figure 7. The near-surface ocean is almost uniform below breaking waves. The velocity gradients beneath the surface are found to be 2 to 5 times weaker than in the wall boundary and the near-surface ocean is almost uniform due to wave-breaking stirring. Mean profiles of (a) magnitude and (b) direction of relative current velocity with respect to a 5-m depth Δu j,6 normalized by wind speed in the wind speed ranges U 23 from 4 to 8 m s 1 (open circles) and from 8 to 12 m s 1 (open squares). Dashed line indicates log layer velocity profile. (After Kudryavtsev et al., 2008.)

Li et al. (2013) simulated the observations collected at the Martha s Vineyard Coastal Observatory s Air-Sea Interaction Tower during the CBLAST experiment in 2003 (Fig. 8) using an LES model. The model showed that breaking waves dominated turbulence generation near the ocean surface over the Stokes term production. In fact, diagnostic analysis of the TKE budget in the model of Li et al. (2013) shows a dominant balance between turbulent transport and dissipation near the surface and a dominant balance between shear production and dissipation at deeper depths. Although the Stokes production is a significant term in the TKE budget balance near the surface, it is smaller than shear production. Langmuir circulation could develop deeper in the water column; though, according to Figure 8, the Stokes production is much smaller than shear production and dissipation term also below the wave stirred layer. It is therefore not obvious that large eddies reproduced by this LES model below the layer of wave breaking turbulence were driven by Stokes drift. Figure 8. Comparison of TKE budget terms including wave breaking and Langmuir circulation (Li et al., 2013). How can Langmuir circulation exist under breaking waves? In this work we consider a new mechanism for creation of Langmuir cells and ramp-like structures, which can help to address this question. Based on the analysis of the turbulence and Stokes production terms, we

have found that in certain conditions the Stokes drift may not be the main driving mechanism for Langmuir circulation. We have completely eliminated the Stokes drift from the LES model but added injection of mixing due to wave breaking in the near surface layer of the ocean (Fig. 9). The model is able to reproduce both Langmuir cells and ramp-like structures coexisting in space though intermittent in time. This model does not require the Stokes drift term and thus is not locked to the wave direction. It may explain the observations similar to those shown in Figure 2 when the Langmuir cell direction does not coincide with the wave propagation direction. Figure 9. Langmuir circulation and ramp-like structures reproduced by accounting for mixing due to wave-breaking turbulence in the near-surface layer of the ocean. Langmuir cells and ramp-like structures coexist in space, but are intermittent in time. The numerical domain is 500 m long, 200 m wide and 80 m deep. The vertical resolution of the mesh near the surface is 0.1 m, gradually increasing with depth. A periodic boundary condition is set along the tank. A 0.1 N m -2 wind stress (indicated by a vector) was applied to the top of the domain and was equivalent to approximately 8 m s -1 wind speed at a 10 m height. The proposed model of Langmuir circulation may compete with that of Craik and Leibovich (1976) within a certain range of depths and wind/wave conditions (in particular, under

high wind speed conditions when the vortex force term including Stokes drift can be suppressed due to strong near-surface turbulence). Discussion A plausible mechanism for the generation of ramp-like structures in the near-surface layer of the ocean has been proposed by Soloviev and Lukas (2014). It is based on the analysis of ramp-like structures in the atmospheric boundary layer over the ground by Smith and Walker (1997) and McNaughton and Brunet (2002). Figure 10. Schematic cross-flow section of the low-speed streaks near the ground (Smith and Walker, 1997). Townsend (1961) divided the near-wall turbulence into the active part (transporting momentum) and the inactive part (not transporting momentum). According to McNaughton and Brunet (2002) inactive motions could initiate active, coherent ejection motions, which carry much of the momentum. These authors concluded that the inactive motions take the form of streak patterns of faster and slower air (Fig. 10), and the streaks are aligned with the surface wind. The narrow convergence lines of uplifted, slower air are formed due to the high-speed streams of subsiding and laterally spreading air. The difference in speeds of the various parts of the flow thus creates convergence zones where the high-speed streams overtake the slower moving streaks. In each of these zones, the faster air stream at first simply passes about the slower streak, creating a zone of strong shear between the faster and slower air streams. The velocity profile along normals to this interface is strongly inflected, forming a classic source of instability in the flow. It initiates a series of transverse roll vortices, just as similar inflections do in plane mixing layers, but here the roll vortices are draped across the spine of the engulfed streak. These vortices describe gentle arcs where the streak is low and board, but become croissant- or horseshoe vortices over taller, more upright parts of streaks. Well-formed horseshoe vortices can then assume a life of their own, continuing to grow by taking vorticity and turbulent kinetic energy (TKE) from the mean flow

itself. The mean shear also rotates these coherent vortices forward until, by a combination of growth and rotation, they contact the ground to form a dam with strong inflows along the ground produced by the rotation of the vortex arms and the main flow presenting pressing in from behind. With nowhere else to go, the trapped air squirts backward and outward into the flow. This squirt is usually described as an ejection/sweep event. (Soloviev and Lukas 2014.) Figure 11a illustrates a schematic of the first ejection formation on a streak. McNaughton and Brunet s mechanism is consistent with a wide range of results from laboratory and atmospheric boundary-layer experiments. (a) (b) Figure 11. (a) The vortices lying across the spine of the streak take on a horseshoe or hairpin shape and can grow to the point where they contact the ground and cause a vigorous ejection of fluid (McNaughton and Brunet, 2002). The hairpin shape initiates a pair of counter rotating vertical vortices when touches the ground. (b) In application to the near-surface layer of the ocean, a pair of counter rotating vertical vortices is formed when the hairpin shape touches the sea surface, from below. Soloviev and Lukas (2014) analyzed the effect of a similar pattern but developing from the water side of the air-sea interface. In the same way, a pair of counter rotating vertical vortices is formed when the hairpin shape touches the sea surface and ejects fluid from the boundary layer (Fig. 11b). The vertical vortices initiate a converging streak via the Craik-Leibovich 2

mechanism (Fig. 12). This mechanism resembles convective initiation of longitudinal rolls in the atmospheric boundary layer (Brown, 1991). Figure 12. The vertical vortices in the near-surface layer of the ocean produced by the hairpin shape initiate a converging streak via the Craik-Leibovich (CL2) mechanism. Diagram adopted from Sullivan et al. (2007), though no Stokes vortex force is required. Conclusions We have developed a concept, which links the Langmuir circulation and ramp-like structures to wave stirring of the near-surface layer of the ocean. This mode of Langmuir circulation does not require Stokes drift and is locked to the wind (but not wave) direction. Using computational fluid dynamics LES model, we have been able to reproduce both Langmuir cells and ramp-like structures coexisting in space though intermittent in time. This mechanism for generation of Langmuir circulation may be explained with the CL2 theory but without the Stokes drift term. This is a coupled mode of Langmuir circulation and ramp-like structures. This mode can dominate over the traditional model within a certain range of wind/wave conditions (in particular, when the waves are not fully developed). The process has some resemblance to the formation of longitudinal rolls in the atmospheric boundary layer.

Acknowledgements We are grateful to Roger Lukas (UH), Mark Donelan (UM), and Eugene Terray (WHOI) for important discussion of the problem. This work has been supported by the Gulf of Mexico Research Initiative Consortium for Advanced Research on the Transport of Hydrocarbons in the Environment (CARTHE). References Antonia, R. A., Chambers, A. J., Friehe, C. A., Van Atta, C. W., 1979: Temperature ramps in the atmospheric surface layer. Journal of Atmospheric Science 36, 99 108. Benilov, A. Y., Ly, L. N., 2002: Modeling of surface waves breaking effects in the ocean upper layer. Mathematical Computational Models 35, 191 213. Brown, R. A., 1991 Fluid Mechanics of the Atmosphere. International Geophysics Series 47, Academic Press, San Diego. Craik, A. D. D., Leibovich, S., 1976: A rational model for Langmuir circulations. Journal of Fluid Mechanics 73, 401 426. Kudryavtsev, V., Shrira, V., Dulov, V., Malinovsky, V., 2008: On the vertical structure of winddriven sea currents. Journal Physical Oceanography 38, 2121-2144. Li, S., Li, M., Gerbi, G. P., Song, J. B., 2013:, Roles of breaking waves and Langmuir circulation in the surface boundary layer of a coastal ocean. Journal of Geophysical Research Oceans 118, 5173 5187. McNaughton, K. G., Brunet, Y., 2002: Townsend s hypothesis, coherent structures and Monin- Obukhov similarity. Bound-Layer Meteorology 102,161 175. Pollard, R.T. 1977: Observations and theories of Langmuir circulations and their role in near surface mixing. In: Angel M (ed) A voyage of discovery: George Deacon 70th anniversary volume, Pergamon Press, Oxford, p 696. Phong-Anant, D., Antonia, R. A., Chamber, A.J., Rajagopalan, S., 1980: Features of the organized motion in the atmospheric surface layer. Journal of Geophysical Research 424 432. Smith, C. R., Walker, J. D. A., 1997: Sustaining mechanisms of turbulent boundary layers: the role of vortex development and interaction. In: Panton RL (ed) Self-sustaining

mechanisms of wall turbulence. Advances in Fluid Mechanics 15, Computational Mechanics Publications, Southampton, pp 273 308. Soloviev, A.V. and Lukas, R., 2014: The Near-Surface Layer of the Ocean: Structure, Dynamics, and Applications (Second edition), Springer, NY Soloviev, A.V., 1990: Coherent structure at the ocean surface in the convectively unstable conditions. Nature 346,157 160. Sullivan, P. P., McWilliams, J. C., Melville, W. K., 2007: Surface gravity wave effects in the oceanic boundary layer: large-eddy simulation with vortex force and stochastic breakers. Journal of Fluid Mechanics 593, 405-452. Thorpe, S. A., Hall, A. J., 1987: Bubble clouds and temperature anomalies in the upper ocean. Nature 328, 48 51. Thorpe, S. A., Jackson, J. F. E., Hall, A. J., Lueck, R. G., 2003: Measurements of turbulence in the upper ocean mixing layer using Autosub. Journal of Physical Oceanography 33, 122 145. Townsend, A., 1961: Equilibrium layers and wall turbulence. Journal of Fluid Mechanics 11, 97 120. Wijesekera, H. W., Paulson, C. A., Huyer, A., 1999: The effect of rainfall on the surface layer during a westerly wind burst in the western equatorial Pacific. Journal of Physical Oceanography 29, 612 632.