1. Introduction
There is a continued interest in improving the prediction and parameterization of momentum, energy, heat, and gas exchanges between the ocean and atmosphere. It is well known that the coupling between wind and ocean surface waves modifies turbulent air–sea fluxes, thereby playing a substantial role in the development of weather and climate patterns forecast by numerical models (Cronin et al. 2019). Yet, many uncertainties remain as to how wind–wave coupling impacts air–sea momentum and scalar fluxes, as well as wave growth and dissipation, under a range of sea states (Sullivan and McWilliams 2010).
Coupled ocean, atmosphere, and wave models often struggle to accurately forecast the evolution of wave fields, ocean currents, storm surge, and weather events without accounting for the effects of wind-waves and swell on the wind stress or drag coefficient Cd (Moon et al. 2003, 2004, 2009; Fan et al. 2009; Donelan et al. 2012; Reichl et al. 2014; Cronin et al. 2019). Most existing models rely on an empirically derived bulk Cd that uses Monin–Obukhov similarity relationships to define the wind stress as a simple function of wind speed in neutral conditions (e.g., Large and Pond 1981; Edson and Fairall 1998). To account for the wave effects (sea state) in a relatively simple manner, many studies have addressed the impacts of wave parameters such as wave age (
A number of more complex models and parameterizations have been developed to account for the effects of wave-driven turbulent processes near the wavy surface under strong wind forcing. Such processes include airflow separation over breaking waves (Kudryavtsev and Makin 2001; Makin and Kudryavtsev 2002; Donelan et al. 2006; Mueller and Veron 2009; Kukulka et al. 2007; Kukulka and Hara 2008; Suzuki et al. 2013), ejection of sea spray and spume from wave crests (Andreas 2004; Kudryavtsev and Makin 2011; Richter and Sullivan 2013; Veron 2015), enhanced bubble production, air entrainment and gas transfer by breaking waves (Deike et al. 2017; Deike and Melville 2018), and near-surface ocean currents driven by waves (Teixeira 2018; Wang et al. 2019), to name a few. A variety of parameterizations have also been developed to represent the effects of wind-wave misalignment on wave growth and dissipation (e.g., Tolman and Chalikov 1996; Meirink et al. 2003; Kudryavtsev and Makin 2004; Ardhuin et al. 2007) and on the surface wind stress and drag coefficient (Bourassa et al. 1999; Grachev et al. 2003; Suzuki et al. 2010; Roekel et al. 2012).
Several studies have developed sea-state-dependent parameterizations of wind stress (or Cd) in complex sea states and extreme winds (Moon et al. 2004; Hara and Belcher 2002, 2004; Fan et al. 2009; Donelan et al. 2012; Reichl et al. 2014). Because field observations in these conditions are limited, the physical mechanisms for the observed leveling off (or reduction) of Cd in extreme winds remain to be fully explained (e.g., Powell et al. 2003; Black et al. 2007; Holthuijsen et al. 2012). Donelan (2004) and Donelan et al. (2006) suggest that under strong wind forcing, airflow separation over steep waves may be an important feature that modifies Cd by causing airflow to skip over troughs and reattach at each crest, rendering the troughs invisible to the wind and reducing the overall sea surface roughness. Following from these studies, Peirson and Garcia (2008) and Grare et al. (2013) use field and laboratory results to emphasize the importance of the wave steepness ak and the pressure wave slope correlation in determining the form drag and wave growth rate. They observe a reduced wave growth rate with increasing ak, and a phase shift of the pressure field which suggests reattachment of separated airflow onto the windward face of the following wave.
Earlier studies have described airflow separation as a process which occurs only after waves have started breaking (Banner and Melville 1976; Banner 1990; Belcher and Hunt 1998; Reul et al. 2008), but recent laboratory studies using particle image velocimetry (PIV) in a wind-wave flume have captured intermittent airflow separation events over both breaking and nonbreaking waves (Veron et al. 2007; Troitskaya et al. 2011; Buckley and Veron 2016, 2019; Savelyev et al. 2020; Yousefi et al. 2020). The PIV results from a recent study by Savelyev et al. (2020) in a wind-wave tank capable of sustaining strongly energetic wave fields (e.g., high ak) show dampened turbulent kinetic energy in the water directly below strongly forced waves possibly due to enhanced airflow separation and reduced Cd. In laboratory observations, intermittent airflow separation events have two main features in common: detachment of a high vorticity shear layer from the steep wave crest, and weak stagnant velocity in a region below (the “dead zone”). In this region, recirculation (closed streamlines) may be present in the phase-averaged flow fields shown in a reference frame moving with the wave phase speed. The recirculation patterns associated with airflow separation are distinct from those present over more mature waves for which a coherent critical layer height (where wind speed is equal to the wave phase speed) is visible above the wavy surface. It should be noted that the zero-wall stress criterion traditionally used to define separation points is not generally applicable when the fluid velocity is not zero at the boundary, e.g., flows over rotating cylinders. Boundary movement following or opposing the overlying airflow modulates separation (e.g., Gad-el Hak and Bushnell 1991).
Numerical turbulence models have been used to gain more insight into the wind turbulence over surface waves. These have included direct numerical simulations (DNS; Sullivan et al. 2000; Yang and Shen 2010, 2017; Yang et al. 2018; Druzhinin et al. 2012; Liu et al. 2010) and large-eddy simulations (LES; Suzuki et al. 2013; Hara and Sullivan 2015; Sullivan et al. 2018a,b; Husain et al. 2019; Hao and Shen 2019; Åkervik and Vartdal 2019; Jiang et al. 2016; Yang et al. 2013; Cao and Shen 2021; Liu et al. 2010), both of which have enabled a detailed exploration of the flow characteristics that may modify the wind stress, wave growth, and dissipation over a wide range of wave parameters (e.g., wave age, wave steepness). Some have been able to reproduce laboratory observations reasonably well (Troitskaya et al. 2014; Sullivan et al. 2018b; Husain et al. 2019). In particular, Husain et al. (2019) have compared LES with the PIV results of Buckley and Veron (2016) for airflow over a train of steep, strongly forced waves in a laboratory wind-wave flume. Both model and laboratory results exhibit a phase-averaged signature of frequent airflow separation. The reasonable validation of LES results against observations by Husain et al. (2019) provides the basis for the current study to use LES to explore the airflow turbulence (occurrence and effects of intermittent airflow separation, in particular) over an extended range of wind and wave conditions expected in the open ocean.
In Part I of this study, we use the identical LES approach to explore a range of wave ages for waves following and opposing wind. A number of previous laboratory studies have addressed wind opposing waves. Young and Sobey (1985) have measured pressure fields to be nearly symmetric about the wave crest, similar to potential flow theory and consistent with previous field observations (Snyder et al. 1981; Hasselmann and Bösenberg 1991). Peirson et al. (2003) and Mitsuyasu and Yoshida (2005) have measured the evolution of waves opposing wind and have found considerable wave decay, consistent with the results of previous numerical simulations using Reynolds-averaged Navier–Stokes (RANS) equations (Al-Zanaidi and Hui 1984; Harris et al. 1995; Mastenbroek 1996; Cohen 1997). Donelan (1999) has also measured wave decay using a pressure–slope correlation, finding that the strong pressure signal in opposing wind can result in substantial form drag despite the absence of a noticeable phase shift in the dominant pressure field. A more recent modeling study of wind opposing waves by Cao et al. (2020) using wall-resolved LES finds a nearly symmetric pressure signal varying with ak and
Despite the existing literature, the physical mechanisms that modify the wind stress, form drag, and wave decay for waves in opposing winds are still not clearly understood. In particular, few studies have explored how waves opposing wind may enhance the effective surface roughness length and the drag coefficient, even though waves opposing wind are quite common and significantly modify the drag coefficient in tropical cyclone conditions (Reichl et al. 2014; Chen and Curcic 2016; Chen et al. 2020). The goal of this study is to expand upon the existing literature to address how opposing wind and waves impact the turbulence in the airflow that modifies these parameters, particularly for steep waves over a range of wave phase speeds, both positive and negative, relative to wind forcing (
2. Methods
a. Large-eddy simulation setup
We use the LES methodology identical to previous studies (Sullivan et al. 2014; Hara and Sullivan 2015; Sullivan et al. 2018b; Husain et al. 2019), which employs a pressure-driven channel flow over a wavy surface propagating through a rectangular domain with doubly periodic horizontal boundaries. This LES study entirely focuses on the airflow and the wave motion is prescribed, i.e., the wave dynamics are decoupled from those of air. We define time t, along-wave x coordinate, cross-wave y coordinate, and vertical coordinate z pointing upward with z = 0 at the mean water surface. Velocities (u, υ, w) are in the (x, y, z) directions.
For most conditions in our study, we consider a linear monochromatic surface wave train with h(x, t) = a cos(kx − ωt), where a is the wave amplitude, k is the wavenumber, ω is the angular frequency, and
The dimensions of the computational domain are lx × ly × lζ, where lx = ly = 5λ and lζ = 2.435λ, with λ = 2π/k as the wavelength. It is discretized with (Nx, Ny, Nζ) = (256, 256, 256) grid points, making the horizontal resolution Δx = Δy = 0.01 953λ. The vertical spacing ratio gradually increases away from the surface so that the ratio between neighboring cells is held constant at 1.0028, with the first point off the water surface located at
b. Wind forcing
The LES setup models an environment similar to that of a wind-wave flume, where an externally imposed horizontal pressure gradient ∂P/∂x is balanced by a surface wind stress such that τs = (∂P/∂x)lζ. Here, the stress and the pressure have been divided by air density ρa and have dimensions of velocity squared. The surface friction velocity is defined as
c. Subgrid-scale and surface roughness parameterizations
Turbulent flow in LES is spatially filtered such that dominant-scale turbulence is resolved while subgrid-scale (SGS) fluxes below a filter threshold are parameterized using a conventional TKE-closure SGS parameterization described in more detail in Moeng (1984), Sullivan et al. (2014), and Moeng and Sullivan (2015), and used by Hara and Sullivan (2015), Sullivan et al. (2018b), Husain et al. (2019), and a number of other studies.
Similar to previous studies, we employ a wall-modeled LES. Along the wavy surface, the local instantaneous tangential stress is parameterized based on the local instantaneous mean wind shear (determined from the difference between the surface tangential water velocity and the tangential wind velocity at the first grid point off the surface) by applying the law of the wall (a log profile) with a prescribed background surface roughness length zob. The parameter zob represents the bulk effect of viscosity (which is more important in laboratory conditions) and subgrid roughness elements such as higher frequency waves (which dominate in open ocean conditions) on the local frictional stress.
In previous studies (Hara and Sullivan 2015; Sullivan et al. 2018b), the normalized background surface roughness has been set at kzob = 2.7 × 10−3 to represent typical strongly forced wind-wave conditions in a laboratory setting. In open ocean conditions with a spectrum of waves, kzob likely varies significantly depending on wind speed, the scale of resolved waves of interest, and other environmental factors such as surfactants. It is also expected that zob, which represents unresolved wave effects, should vary with the phase of the resolved waves because shorter waves are known to be modulated by longer waves (e.g., Gent and Taylor 1976; Gent 1977; Kudryavtsev and Makin 2002).
In our recent study (Husain et al. 2019) we have systematically investigated how varying kzob affects the airflow turbulence over steep waves (ak = 0.27) in strongly forced conditions. The results show that the flow field is quite sensitive to the zob value specified near the crest; a higher zob increases the frequency of intermittent airflow separation events and enhances the resulting signature in the phase averaged flow fields. However, the airflow is hardly affected if the zob value is altered away from the crest. In the same study, the effect of wave-phase-dependent surface drift velocity has been investigated as well. The results indicate that the drift velocity added near the wave crest simply increases the wind speed by the same amount everywhere without affecting the airflow turbulence characteristics, and that the drift velocity added away from the crest has very little impact.
Since the effects of varying zob and surface drift have been investigated previously, in this study we keep the normalized background surface roughness held constant at kzob = 2.7 × 10−3 and impose zero surface drift velocity for all simulations; that is, we do not repeat the sensitivity study of varying roughness length and drift velocity, acknowledging that their impacts are potentially important. We then focus on investigating the effects of varying wind forcing
Note that since our logarithmic wall model is based on the assumption of turbulence in equilibrium, its applicability may be questionable where flow separation occurs. However, our previous study (Husain et al. 2019) shows that changing the roughness value below flow separation areas (away from the wave crest) in strongly forced cases has little impact on the results, suggesting that our LES results are not very sensitive to the wall modeling in such areas.
d. Simulations
While the actual LES of waves opposing wind is performed by reversing the wind direction as described above, in the following sections we differentiate waves following wind and waves opposing wind by the sign of the wave phase speed c; that is, the wind always blows in the positive x direction (wind stress τs is always negative), and the waves propagate in the positive/negative x direction for following/opposing cases.
In total, we perform simulations for five wave ages in following wind (
Each simulation is run for approximately 130 000 time steps and averaged over the last 60 000 time steps after the wind field has reached a statistically steady state. Sullivan et al. (2014) and Sullivan et al. (2018b) provide a full description of the LES algorithm and numerical methods used to solve the governing equations. See Table 1 for more details on the conditions simulated in this study.
List of run conditions and results of roughness enhancement zo/zob and nondimensional wave growth/decay coefficient cβ for 11 LES simulations. The letters “f” and “o” in the run name represent waves following and opposing wind, respectively. The letter “s” in the run name represents Stokes waves. Nondimensional parameters used in LES cases include wave age (
e. Data analysis
3. Results and discussion
a. Two-dimensional phase averaged airflow above waves following wind
In this section all flow fields presented are normalized by
(left three columns) Normalized phase-averaged flow fields in the ξ–z coordinate and (right three columns) the mapped ξ–ζ coordinate for (top three rows) waves following wind (
Citation: Journal of Physical Oceanography 52, 1; 10.1175/JPO-D-21-0043.1
(left) Normalized phase-averaged flow fields in the ξ–z coordinate and (center) the mapped ξ–ζ coordinate for (top three rows) waves following wind (
Citation: Journal of Physical Oceanography 52, 1; 10.1175/JPO-D-21-0043.1
(left three columns) Normalized phase-averaged flow fields in the ξ–z coordinate and (right three columns) the mapped ξ–ζ coordinate for (top three rows) waves following wind (
Citation: Journal of Physical Oceanography 52, 1; 10.1175/JPO-D-21-0043.1
First, we examine the cases of waves following wind (top three panels of Figs. 1–3). In the case of strongly forced waves at
Buckley and Veron (2016) have been able to demonstrate with high-resolution particle image velocimetry (PIV) that near-surface wind flow patterns are modified by transient, sporadic detachment of airflow from the crests of steep, young waves. Instantaneous measurements of u and w over laboratory waves capture a layer of enhanced spanwise vorticity developing and ejecting away from the wave crest, resulting in weak, stagnant, sometimes negative airflow in the trough (see Buckley and Veron 2016, their Fig. 6). Previous LES results (e.g., Hara and Sullivan 2015; Sullivan et al. 2018b), and a more recent comparison between these observations and LES by Husain et al. (2019), have also found that intermittent airflow separation events can occur frequently over steep, nonbreaking waves, and their phase-averaged character in observations can be well reproduced in LES. This suggests that while these separation events are not persistent in either the observations or LES, they occur frequently enough in steep, strongly forced conditions to leave a signature in the phase-averaged flow fields.
These intermittent airflow separation events also affect the phase-averaged pressure field (Figs. 2c,C,G). It is well known that the pressure–wave slope correlation determines the form drag and the wave growth rate (e.g., Peirson and Garcia 2008; Grare et al. 2013), but the role of airflow dynamics in shifting the phase and magnitude of maximum surface pressure is still not well understood. The LES results of Sullivan et al. (2018b) associate the location and magnitude of maximum pressure with reattachment of detached flow onto the windward face of the following wave. Our results also suggest that the reattachment of the separated flow influences the location of the high pressure on the windward face of the wave (Figs. 2c,C,G) and the resulting form drag and wave growth rate (discussed in section 3g).
As the wave age (
Unlike the slow wave case (
b. Two-dimensional phase averaged airflow above waves opposing wind
Next, we examine the cases of waves opposing wind (bottom panels in Figs. 1–3). One immediately notices that the flow field is dominated by a strong pressure perturbation (Figs. 2d–f,D–F) in phase with −h (negative of the wave elevation), and a strong vertical velocity perturbation (Figs. 1p–r,P–R) in phase with ∂h/∂x (the wave slope). This result is consistent with the wall-resolved LES results of Cao et al. (2020) for waves opposing wind, as well as previous laboratory and field measurements (Snyder et al. 1981; Young and Sobey 1985; Hasselmann and Bösenberg 1991). Cao et al. (2020) show that these flow features are well explained by a simple linear inviscid model. This trend of increasingly stronger along-wave pressure gradient is reminiscent of potential flow, as observed by Young and Sobey (1985). However, the pressure field in phase of −h does not contribute to the wave growth rate or the form drag; only the small out-of-phase component does. Cao et al. (2020) discuss how the strong turbulence very near the surface plays an important role in determining the magnitude of this component.
Similarly, our results show that the near surface turbulence fields contain strong wave induced perturbations. The most notable feature near the wavy surface is that there is a clear progression in the flow structure from positive to negative wave ages (more specifically, from large positive to small positive to small negative to large negative wave ages; that is, from top to bottom in Figs. 1–3). In particular, the flow fields of TKE, ϵ, and ωy are surprisingly similar between the cases of
As the wave speed increases from
One surprising feature is that
Although the pressure field appears dominated by the perturbation in phase with −h, a significant out-of-phase component exists which is not apparent in the flow fields, but nonetheless results in an increase of the form drag and wave decay rate as
c. Instantaneous airflow features
To demonstrate the transient character of the flow field over a range of
Normalized instantaneous vorticity fields [
Citation: Journal of Physical Oceanography 52, 1; 10.1175/JPO-D-21-0043.1
With faster waves at
In the case of slow waves opposing wind (
d. Vertical profiles of horizontally averaged wind fields
In the following subsections we investigate the vertical profiles (dependence on ζ) of wind variables averaged horizontally in the mapped coordinate (mean wind speed, mean wind shear, mean TKE, and the terms in the momentum and energy budget equations) as well as the enhancement of the equivalent surface roughness due to waves and the wave growth (decay) rates for waves following (opposing) wind. In particular, we attempt to explain how these quantities are affected by the physical mechanisms identified in the two-dimensional flow analysis in the previous subsections. All the profiles are displayed up to kζ = 4 because the results above this elevation are affected by the reduced wind stress and the LES top boundary.
Figures 5a–f display normalized horizontal mean profiles of wind speed (
Normalized vertical profiles of horizontally averaged wind speed (
Citation: Journal of Physical Oceanography 52, 1; 10.1175/JPO-D-21-0043.1
For waves following wind (Fig. 5a), the far field wind profiles above the wave boundary layer (above about kζ = 1) are roughly parallel to, but shifted to the left of, the wind profile over a flat surface with a nondimensional background surface roughness of kzob = 2.7 × 10−3 (gray lines). Here, the solid gray line represents the wind profile modified by the linearly decreasing wind stress in kζ, and the dashed gray line is the wind profile for constant stress in kζ. Since the background roughness length (accounting for the form drag of unresolved small waves and the viscous stress) along the wavy surface is identically set at kzob = 2.7 × 10−3 in all simulations, the decrease of the far field wind speed indicates that the waves enhance the effective roughness length zo (determined by extrapolating the wind profile above the wave boundary layer toward the surface) relative to the background roughness length zob.
Specifically, for each case we roughly estimate zo by horizontally shifting the flat wall wind profile (gray solid line) to match the wind speed profile above the wave boundary layer (matching the wind speed at kζ = 4 for simplicity), then we find the height where the shifted flat wall wind speed becomes zero. We find that the slowest waves (
The mean normalized wind shear is shown in Fig. 5c. Similar to the gray lines for mean wind speed in Fig. 5a, gray lines here represent the mean wind shear profile unmodified by waves for linearly decreasing wind stress in kζ (solid) and constant wind stress in kζ (dashed). When the wind shear profile deviates to the right (left) of the gray solid line, the wind shear is enhanced (reduced) due to the wave effect. Since the mean wind speed must approach zero at the background roughness height in all simulations (z = zob), the shift of the far field wind profile relative to the flat wall profile means that the mean wind shear is modified by waves inside the wave boundary layer. Here, the normalized mean wind shear is defined and plotted such that the area integral of its deviation from the flat wall case is approximately proportional to the deviation of the normalized far field mean wind speed from the flat wall profile (see Fig. 2 and discussion in Hara and Sullivan 2015).
For waves following wind, the wave age plays a significant role in the character of the mean wind shear. Slower waves (
As the wave age increases to
Note that the normalized mean shear remains slightly above the flat wall case (gray solid line) at the top of the domain (kζ = 4) for the cases of
The mean normalized 3D TKE profiles are shown in Fig. 5e. For the two slow wave cases the enhancement of TKE is generally located at about the same elevation as the enhancement of the mean wind shear in Fig. 5c. As
Next, we examine the case for waves opposing wind. With slow waves at
As the wave speed increases and
Similar to the case of faster waves following wind
e. Momentum budget in mapped coordinate
The horizontally averaged momentum budget (or wind stress partition) as described in Eq. (7) is shown in Figs. 6A,C (upper panels) for
(left) Normalized phase-averaged fields of wave-coherent stress (
Citation: Journal of Physical Oceanography 52, 1; 10.1175/JPO-D-21-0043.1
The two-dimensional fields of the phase-averaged normalized turbulent stress
For the slow waves following wind (
As the wave age increases to
For slow waves opposing wind at
As the wave speed increases and
The two-dimensional fields of
f. Energy budget and turbulence closure parameterization
In Figs. 7a,b, the second, third, and fourth terms of Eq. (10) are plotted in solid, dot–dashed, and dotted lines with colors corresponding to their respective
(a),(b) Normalized vertical profiles of horizontally averaged energy budget terms. The first (pressure gradient), second (shear production), third (transport) and fourth (dissipation) terms of Eq. (10) are solid gray, solid, dot–dashed, and dotted lines, respectively, with thin dotted lines near zero equaling the sum of all energy budget terms. (c),(d) Normalized vertical profiles of horizontally averaged turbulent stress (
Citation: Journal of Physical Oceanography 52, 1; 10.1175/JPO-D-21-0043.1
Previous modeling studies of the vertical mean wind profile and the drag coefficient over a surface wave train (or a spectrum of waves, i.e., many surface wave trains superimposed) have sought to close the turbulence in the wave boundary layer by parameterizing the eddy viscosity (K) or the TKE dissipation rate (〈ϵ/J〉) using the turbulent stress 〈τt〉 that varies with height due to the wave influence (e.g., Makin and Kudryavtsev 1999; Hara and Belcher 2004). Thus, one area of interest in the present study is to determine whether the character of the mean normalized wind shear (equivalent to the mean shear production term of the energy budget, solid lines in Figs. 7a,b and identical to Figs. 5c,d), the mean dissipation [last term in Eq. (10), dotted lines in Figs. 7a,b], and the mean turbulent stress share similarities in character over a range of wind-wave conditions. For this reason, we have included the profiles of the mean turbulent stress in Figs. 7c,d (previously shown in Figs. 6C,D), and the profiles of the normalized eddy viscosity
In all cases, the shear production roughly balances the viscous dissipation throughout the wave boundary layer, with a relatively modest contribution from the transport term (except for
One notable exception to the generally good correlation between the dissipation rate and the turbulent stress is that for the cases of waves opposing wind, the shear production and viscous dissipation are both significantly reduced (by as much as 1/2) above kζ = 1 all the way to the top (kζ = 4), even if the turbulent stress 〈τt〉 is almost equal to the total wind stress 〈τtot〉 (i.e., the wave effect on the turbulent stress is negligible) in the same kζ range. As discussed earlier, this reduced mean wind shear makes a significant contribution to the enhancement of the equivalent roughness length zo and the drag coefficient. A turbulence closure model based on the wave modified 〈τt〉 alone would completely miss this impact.
The profiles of the normalized eddy viscosity
g. Wave growth/decay rate and equivalent roughness length
The normalized phase-dependent surface stress distribution is plotted in Figs. 2G,H for waves following wind, and in Figs. 2I,J for waves opposing wind. Since the total normal stress is a sum of the pressure and the turbulent normal stress, both the total normal stress (solid line) and the pressure stress alone (dotted line) are shown in Figs. 2G,I. The turbulent tangential surface stress is presented in Figs. 2H,J.
(top) Wave growth/decay coefficient |cβ| for waves following wind (black lines) and for waves opposing wind (red lines) as a function of
Citation: Journal of Physical Oceanography 52, 1; 10.1175/JPO-D-21-0043.1
Consistent with Husain et al. (2019) for waves with
For waves opposing wind, as
At
In Fig. 9, our estimated cβtot and cβp values (Fig. 9a) and β/ω values (Figs. 9b,c) for waves opposing wind are compared to the previous LES study of Cao et al. (2020) as well as the results of Harris et al. (1995), Cohen (1997), Donelan (1999), Peirson et al. (2003), and Mitsuyasu and Yoshida (2005). Since most previous studies focused on wave attenuation due to pressure only, they should be compared with our results for cβp or corresponding β/ω (thin red lines).
(top) Comparison of wave decay coefficient |cβ| as a function of wave age
Citation: Journal of Physical Oceanography 52, 1; 10.1175/JPO-D-21-0043.1
Our results, plotted against
In Fig. 8b we summarize the estimated values for zo/zob, which represents the enhancement of the effective roughness length zo due to resolved waves relative to the background roughness length zob. The results of zo/zob are shown in a log scale because the increasing wind speed above the wave boundary layer due to resolved waves is proportional to log(zo/zob). The figure highlights the strong dependence of the effective roughness length and the drag coefficient on
It is interesting that while the magnitude of cβp is similar (i.e., the pressure form drag is similar) between the cases of
Hara and Belcher (2004) show that inside the constant stress layer the downward energy flux at the top of the wave boundary layer is roughly equal to the mean wind speed multiplied by the wind stress
Hara and Belcher (2004) then assume that the reduction of the viscous dissipation inside the wave boundary layer (compared to that over a flat surface) is correlated with the reduction of the turbulent stress, which is caused by the pressure form drag. Therefore, if the pressure form drag is similar between the cases of
In fact, a quick estimate of the difference in the far field wind speed
4. Summary
In this study, we use large-eddy simulation (LES) to simulate turbulent wind flow over steep waves (ak = 0.27) following and opposing the wind for a range of wave speeds relative to wind forcing (
Wind flow over opposing waves results in a strong wave-induced flow perturbation that intensifies and is compressed near the surface as the phase speed of the waves increases. We observe a number of phase-averaged flow features similar to those over slow waves following wind, e.g., enhanced TKE, dissipation, and detached vorticity near the wave crest, as well as reduced TKE, dissipation, and vorticity in the wave trough below the detached enhanced vorticity layer (Figs. 3D–F,J–L,P–R). However, the strong positive wind along the wave shape over opposing waves (in a frame of reference moving with the wave) inhibits apparent separation-like flow patterns. Increases in opposing wave speed intensify the in-phase component of the pressure field (Figs. 2d–f,D–F,I) and make the flow appear to follow the potential wave theory. They also induce a significant out-of-phase component of the pressure field responsible for an increase in the effective surface roughness and wave decay rate (Fig. 8). Our estimated wave decay rates are consistent with those of previous studies, including a recent study using wall-resolved LES (Cao et al. 2020), model studies using RANS solutions (Al-Zanaidi and Hui 1984; Harris et al. 1995; Mastenbroek 1996; Cohen 1997), and laboratory studies (Donelan 1999; Peirson et al. 2003; Mitsuyasu and Yoshida 2005).
It is noteworthy that the observed separation-like signatures of wind over the wave crest are qualitatively similar to flow separations of wind blowing over a rotating cylinder (placed horizontally with its axis perpendicular to the wind), which were investigated by Gad-el Hak and Bushnell (1991) and Degani et al. (1998). This is not surprising because the wind velocity at the wave crest is not zero but positive (negative) for waves opposing (following) wind in a reference frame moving with the wave, and the wind velocity at the top of the rotating cylinder is also positive (negative) if the cylinder rotates forward (backward). Gad-el Hak and Bushnell (1991) notes that there is a close relationship between steady flow over a moving wall and unsteady flow over a fixed wall, and that in these conditions separation points may be lifted above the surface and the traditional criterion of zero surface shear stress does not apply. In fact, they predict that near-surface wind shear becomes negative (positive) over a cylinder rotating forward (backward), which corresponds to a crest of waves opposing (following) wind. This prediction is consistent with our LES results discussed earlier; in particular, producing strong negative vorticity in the trough for faster waves (Figs. 3P–R; also see Fig. 1 in Degani et al. 1998).
Our estimates of the equivalent surface roughness zo (including the effect of resolved waves) relative to the background roughness zob (representing the form drag of unresolved waves and viscosity) show that the enhancement zo/zob is significant for slow waves following wind but decreases as the wave age increases. On the other hand, for waves opposing wind, zo/zob rapidly increases as the wave speed increases. By comparing the results of the slowest waves for both following and opposing wind (
Waves opposing wind often appear when the wind field rapidly changes in space and/or time, a situation commonly encountered under tropical cyclones. Previous modeling efforts of the sea-state-dependent drag coefficient have predicted waves opposing wind may significantly enhance the drag coefficient in such conditions because of the assumed large form drag (Reichl et al. 2014; Chen and Curcic 2016; Chen et al. 2020). Results from the present study provide credible support for such modeling efforts. In addition, our energy budget analysis (section 3f) and the discussion on the roughness length (section 3g) identify the strengths and weaknesses of existing models of mean wind profile and drag coefficient over a spectrum of waves.
In this study we do not propose a new parameterization of the drag coefficient as a function of wave age or other simple wave parameters. This is mainly because the total wind stress is expected to be dependent on integration of the wave form drag due to waves of all scales and directions (e.g., Donelan et al. 2012; Reichl et al. 2014). Only when the entire wave spectrum can be characterized by simple wave parameters (e.g., fetch-dependent growing wind seas under steady uniform wind) can a simple drag coefficient parameterization be feasible. Since waves opposing wind appear when the wind field rapidly changes in space/time, it is unlikely that a simple drag coefficient parameterization is applicable in such conditions. Instead, the aim of this study is to advance our understanding of how waves opposing wind interact with wind, how large the wave decay rate and the wave form drag are, and how the mean wind profile is modified by such waves.
Previous studies suggest that the wave growth/decay rate and the effective roughness length may be significantly modified by sea spray (e.g., Bell et al. 2012; Innocentini and Goncalves 2010). In addition, they may be further modified by the nonlinearity of surface waves (e.g., Zdyrski and Feddersen 2020). In this study the spray effects have not been addressed and the wave nonlinearity effect has been investigated by one simulation only using the second-order Stokes waves. It is certainly desirable to incorporate these effects more fully in future LES studies, particularly because waves opposing wind are common under tropical cyclones, and spray and nonlinear effects may dominate in such high wind conditions. However, there are many conditions where our results are more likely relevant. For example, even under a tropical cyclone there is a large area (away from the eyewall) where wind speed is modest and opposing dominant swell waves are not very steep, with
Acknowledgments.
We acknowledge support of the National Science Foundation (Physical Oceanography) Grant OCE-1458984 (URI). We also acknowledge high-performance computing support from Yellowstone and Cheyenne (doi:10.5065/D6RX99HX) provided by NCAR’s Computational and Information Systems Laboratory, sponsored by the National Science Foundation.
Data availability statement.
All large-eddy simulation data created and used during this study are openly available at http://dx.doi.org/10.17632/8vj68sr4rx.1.
REFERENCES
Åkervik, E., and M. Vartdal, 2019: The role of wave kinematics in turbulent flow over waves. J. Fluid Mech., 880, 890–915, https://doi.org/10.1017/jfm.2019.708.
Al-Zanaidi, M. A., and W. H. Hui, 1984: Turbulent airflow over water waves-a numerical study. J. Fluid Mech., 148, 225–246, https://doi.org/10.1017/S0022112084002329.
Andreas, E. L., 2004: Spray stress revisited. J. Phys. Oceanogr., 34, 1429–1440 https://doi.org/10.1175/1520-0485(2004)034<1429:SSR>2.0.CO;2.
Ardhuin, F., T. H. C. Herbers, K. P. Watts, G. P. van Vledder, R. Jensen, and H. C. Graber, 2007: Swell and slanting-fetch effects on wind wave growth. J. Phys. Oceanogr., 37, 908–931, https://doi.org/10.1175/JPO3039.1.
Banner, M. L., 1990: The influence of wave breaking on the surface pressure distribution in wind–wave interactions. J. Fluid Mech., 211, 463–495, https://doi.org/10.1017/S0022112090001653.
Banner, M. L., and W. K. Melville, 1976: On the separation of air flow over water waves. J. Fluid Mech., 77, 825–842, https://doi.org/10.1017/S0022112076002905.
Belcher, S. E., and J. C. R. Hunt, 1993: Turbulent shear flow over slowly moving waves. J. Fluid Mech., 251, 109, https://doi.org/10.1017/S0022112093003350.
Belcher, S. E., and J. C. R. Hunt, 1998: Turbulent flow over hills and waves. Annu. Rev. Fluid Mech., 30, 507–538, https://doi.org/10.1146/annurev.fluid.30.1.507.
Belcher, S. E., T. M. J. Newley, and J. C. R. Hunt, 1993: The drag on an undulating surface induced by the flow of a turbulent boundary layer. J. Fluid Mech., 249, 557–596, https://doi.org/10.1017/S0022112093001296.
Bell, M. M., M. T. Montgomery, and K. A. Emanuel, 2012: Air–sea enthalpy and momentum exchange at major hurricane wind speeds observed during CBLAST. J. Atmos. Sci., 69, 3197–3222, https://doi.org/10.1175/JAS-D-11-0276.1.
Black, P. G., and Coauthors, 2007: Air–sea exchange in hurricanes: Synthesis of observations from the coupled boundary layer air–sea transfer experiment. Bull. Amer. Meteor. Soc., 88, 357–374, https://doi.org/10.1175/BAMS-88-3-357.
Bourassa, M. A., D. G. Vincent, and W. L. Wood, 1999: A flux parameterization including the effects of capillary waves and sea state. J. Atmos. Sci., 56, 1123–1139, https://doi.org/10.1175/1520-0469(1999)056<1123:AFPITE>2.0.CO;2.
Buckley, M. P., and F. Veron, 2016: Structure of the airflow above surface waves. J. Phys. Oceanogr., 46, 1377–1397, https://doi.org/10.1175/JPO-D-15-0135.1.
Buckley, M. P., and F. Veron, 2019: The turbulent airflow over wind generated surface waves. Eur. J. Mech., 73B, 132–143, https://doi.org/10.1016/j.euromechflu.2018.04.003.
Cao, T., and L. Shen, 2021: A numerical and theoretical study of wind over fast-propagating water waves. J. Fluid Mech., 919, A38, https://doi.org/10.1017/jfm.2021.416.
Cao, T., B.-Q. Deng, and L. Shen, 2020: A simulation-based mechanistic study of turbulent wind blowing over opposing water waves. J. Fluid Mech., 901, A07, https://doi.org/10.1017/jfm.2020.591.
Chalikov, D., and S. Rainchik, 2011: Coupled numerical modelling of wind and waves and the theory of the wave boundary layer. Bound.-Layer Meteor., 138, 1–41, https://doi.org/10.1007/s10546-010-9543-7.
Chen, S. S., and M. Curcic, 2016: Ocean surface waves in Hurricane Ike (2008) and Superstorm Sandy (2012): Coupled model predictions and observations. Ocean Modell., 103, 161–176, https://doi.org/10.1016/j.ocemod.2015.08.005.
Chen, X., I. Ginis, and T. Hara, 2020: Impact of shoaling ocean surface waves on wind stress and drag coefficient in coastal waters: 2. Tropical cyclones. J. Geophys. Res. Oceans, 125, e2020JC016223, https://doi.org/10.1029/2020JC016223.
Cohen, J. E., 1997: Theory of turbulent wind over fast and slow waves. Ph.D. thesis, University of Cambridge, 228 pp., https://doi.org/10.17863/CAM.31086.
Cronin, M. F., and Coauthors, 2019: Air-sea fluxes with a focus on heat and momentum. Front. Mar. Sci., 6, 430, https://doi.org/10.3389/fmars.2019.00430.
Degani, A. T., J. D. A. Walker, and F. T. Smith, 1998: Unsteady separation past moving surfaces. J. Fluid Mech., 375, 1–38, https://doi.org/10.1017/S0022112098001839.
Deike, L., and W. K. Melville, 2018: Gas transfer by breaking waves. Geophys. Res. Lett., 45, 10 482–10 492, https://doi.org/10.1029/2018GL078758.
Deike, L., L. Lenain, and W. K. Melville, 2017: Air entrainment by breaking waves. Geophys. Res. Lett., 44, 3779–3787, https://doi.org/10.1002/2017GL072883.
Donelan, M. A., 1999: Wind-induced growth and attenuation of laboratory waves. Wind-over-Wave Couplings: Perspectives and Prospects, S. G. Sajjadi, N. H. Thomas, and J. C. R. Hunt, Eds., Clarendon, 183–194.
Donelan, M. A., 2004: On the limiting aerodynamic roughness of the ocean in very strong winds. Geophys. Res. Lett., 31, L18306, https://doi.org/10.1029/2004GL019460.
Donelan, M. A., F. W. Dobson, S. D. Smith, and R. J. Anderson, 1993: On the dependence of sea surface roughness on wave development. J. Phys. Oceanogr., 23, 2143–2149, https://doi.org/10.1175/1520-0485(1993)023<2143:OTDOSS>2.0.CO;2.
Donelan, M. A., A. V. Babanin, I. R. Young, and M. L. Banner, 2006: Wave-follower field measurements of the wind-input spectral function. Part II: Parameterization of the wind input. J. Phys. Oceanogr., 36, 1672–1689, https://doi.org/10.1175/JPO2933.1.
Donelan, M. A., M. Curcic, S. S. Chen, and A. K. Magnusson, 2012: Modeling waves and wind stress. J. Geophys. Res., 117, C00J23, https://doi.org/10.1029/2011JC007787.
Druzhinin, O. A., Y. I. Troitskaya, and S. S. Zilitinkevich, 2012: Direct numerical simulation of a turbulent wind over a wavy water surface. J. Geophys. Res., 117, C00J05, https://doi.org/10.1029/2011JC007789.
Edson, J. B., and C. W. Fairall, 1998: Similarity relationships in the marine atmospheric surface layer for terms in the TKE and scalar variance budgets. J. Atmos. Sci., 55, 2311–2328, https://doi.org/10.1175/1520-0469(1998)055<2311:SRITMA>2.0.CO;2.
Edson, J. B., and Coauthors, 2013: On the exchange of momentum over the open ocean. J. Phys. Oceanogr., 43, 1589–1610, https://doi.org/10.1175/JPO-D-12-0173.1.
Fairall, C. W., E. F. Bradley, J. E. Hare, A. A. Grachev, and J. B. Edson, 2003: Bulk parameterization of air–sea fluxes: Updates and verification for the COARE algorithm. J. Climate, 16, 571–591, https://doi.org/10.1175/1520-0442(2003)016<0571:BPOASF>2.0.CO;2.
Fan, Y., I. Ginis, T. Hara, C. W. Wright, and E. J. Walsh, 2009: Numerical simulations and observations of surface wave fields under an extreme tropical cyclone. J. Phys. Oceanogr., 39, 2097–2116, https://doi.org/10.1175/2009JPO4224.1.
Gad-el Hak, M., and D. M. Bushnell, 1991: Separation control. J. Fluids Eng., 113, 5–30, https://doi.org/10.1115/1.2926497.
Gent, P. R., 1977: A numerical model of the air flow above water waves. Part 2. J. Fluid Mech., 82, 349–369, https://doi.org/10.1017/S0022112077000706.
Gent, P. R., and P. A. Taylor, 1976: A numerical model of the air flow above water waves. J. Fluid Mech., 77, 105–128, https://doi.org/10.1017/S0022112076001158.
Grachev, A. A., C. W. Fairall, J. E. Hare, J. B. Edson, and S. D. Miller, 2003: Wind stress vector over ocean waves. J. Phys. Oceanogr., 33, 2408–2429, https://doi.org/10.1175/1520-0485(2003)033<2408:WSVOOW>2.0.CO;2.
Grare, L., W. L. Peirson, H. Branger, J. W. Walker, J.-P. Giovanangeli, and V. Makin, 2013: Growth and dissipation of wind-forced, deep-water waves. J. Fluid Mech., 722, 5–50, https://doi.org/10.1017/jfm.2013.88.
Hao, X., and L. Shen, 2019: Wind-wave coupling study using LES of wind and phase-resolved simulation of nonlinear waves. J. Fluid Mech., 874, 391–425, https://doi.org/10.1017/jfm.2019.444.
Hara, T., and S. E. Belcher, 2002: Wind forcing in the equilibrium range of wind-wave spectra. J. Fluid Mech., 470, 223–245, https://doi.org/10.1017/S0022112002001945.
Hara, T., and S. E. Belcher, 2004: Wind profile and drag coefficient over mature ocean surface wave spectra. J. Phys. Oceanogr., 34, 2345–2358, https://doi.org/10.1175/JPO2633.1.
Hara, T., and P. Sullivan, 2015: Wave boundary layer turbulence over surface waves in a strongly forced condition. J. Phys. Oceanogr., 45, 868–883, https://doi.org/10.1175/JPO-D-14-0116.1.
Harris, J., I. Fultion, and R. Street, 1995: Decay of waves in an adverse wind. Proceedings of the Sixth Asian Congress of Fluid Mechanics, Y. Chew and C. Tso, Eds., Nanyang Technological University, 224–227.
Hasselmann, D., and J. Bösenberg, 1991: Field measurements of wave-induced pressure over wind-sea and swell. J. Fluid Mech., 230, 391–428, https://doi.org/10.1017/S0022112091000848.
Holthuijsen, L. H., M. D. Powell, and J. D. Pietrzak, 2012: Wind and waves in extreme hurricanes. J. Geophys. Res., 117, C09003, https://doi.org/10.1029/2012JC007983.
Husain, N. T., T. Hara, M. P. Buckley, K. Yousefi, F. Veron, and P. P. Sullivan, 2019: Boundary layer turbulence over surface waves in a strongly forced condition: LES and observation. J. Phys. Oceanogr., 49, 1997–2015, https://doi.org/10.1175/JPO-D-19-0070.1.
Husain, N. T., T. Hara, and P. P. Sullivan, 2021: Wind turbulence over misaligned surface waves and air–sea momentum flux. Part II: Waves in oblique wind. J. Phys. Oceanogr., 52, 141–159, https://doi.org/10.1175/JPO-D-21-0044.1.
Innocentini, V., and I. A. Goncalves, 2010: A simple spume droplets and wave stress parameterizations to study the impact on maritime near-surface variables. Geophysical Research Abstracts, Vol. 12, Abstract EGU2010-14701, https://meetingorganizer.copernicus.org/EGU2010/EGU2010-14701.pdf.
Jiang, Q., P. P. Sullivan, S. Wang, J. Doyle, and L. Vincent, 2016: Impact of swell on air–sea momentum flux and marine boundary layer under low-wind conditions. J. Atmos. Sci., 73, 2683–2697, https://doi.org/10.1175/JAS-D-15-0200.1.
Kudryavtsev, V. N., and V. K. Makin, 2001: The impact of air-flow separation on the drag of the sea surface. Bound.-Layer Meteor., 98, 155–171, https://doi.org/10.1023/A:1018719917275.
Kudryavtsev, V. N., and V. K. Makin, 2002: Coupled dynamics of short waves and the airflow over long surface waves. J. Geophys. Res., 107, 3209, https://doi.org/10.1029/2001JC001251.
Kudryavtsev, V. N., and V. K. Makin, 2004: Impact of swell on the marine atmospheric boundary layer. J. Phys. Oceanogr., 34, 934–949, https://doi.org/10.1175/1520-0485(2004)034<0934:IOSOTM>2.0.CO;2.
Kudryavtsev, V. N., and V. K. Makin, 2011: Impact of ocean spray on the dynamics of the marine atmospheric boundary layer. Bound.-Layer Meteor., 140, 383–410, https://doi.org/10.1007/s10546-011-9624-2.
Kukulka, T., and T. Hara, 2008: The effect of breaking waves on a coupled model of wind and ocean surface waves. Part II: Growing seas. J. Phys. Oceanogr., 38, 2164–2184, https://doi.org/10.1175/2008JPO3962.1.
Kukulka, T., T. Hara, and S. E. Belcher, 2007: A model of the air–sea momentum flux and breaking-wave distribution for strongly forced wind waves. J. Phys. Oceanogr., 37, 1811–1828, https://doi.org/10.1175/JPO3084.1.
Large, W. G., and S. Pond, 1981: Open ocean momentum flux measurements in moderate to strong winds. J. Phys. Oceanogr., 11, 324–336, https://doi.org/10.1175/1520-0485(1981)011<0324:OOMFMI>2.0.CO;2.
Liu, Y., D. Yang, X. Guo, and L. Shen, 2010: Numerical study of pressure forcing of wind on dynamically evolving water waves. Phys. Fluids, 22, 041704, https://doi.org/10.1063/1.3414832.
Makin, V. K., and V. N. Kudryavtsev, 1999: Coupled sea surface-atmosphere model: 1. Wind over waves coupling. J. Geophys. Res., 104, 7613–7623, https://doi.org/10.1029/1999JC900006.
Makin, V. K., and V. N. Kudryavtsev, 2002: Impact of dominant waves on sea drag. Bound.-Layer Meteor., 103, 83–99, https://doi.org/10.1023/A:1014591222717.
Mastenbroek, C., 1996: Wind-wave interaction. Ph.D. thesis, Delft Technical University, 119 pp., http://resolver.tudelft.nl/uuid:b91093af-ccae-4d85-894e-af128cc9f59c.
Meirink, J. F., V. K. Makin, and V. N. Kudryavtsev, 2003: Note on the growth rate of water waves propagating at an arbitrary angle to the wind. Bound.-Layer Meteor., 106, 171–183, https://doi.org/10.1023/A:1020835211837.
Mitsuyasu, H., and Y. Yoshida, 2005: Air-sea interactions under the existence of opposing swell. J. Oceanogr., 61, 141–154, https://doi.org/10.1007/s10872-005-0027-1.
Moeng, C.-H., 1984: A large-eddy-simulation model for the study of planetary boundary-layer turbulence. J. Atmos. Sci., 41, 2052–2062, https://doi.org/10.1175/1520-0469(1984)041<2052:ALESMF>2.0.CO;2.
Moeng, C.-H., and P. Sullivan, 2015: Large eddy simulation. Encyclopedia of Atmospheric Sciences, Vol. 4, 2nd ed. G. North, F. Zhang, and J. Pyle, Eds., Academic Press, 232–240, https://doi.org/10.1016/B978-0-12-382225-3.00201-2.
Moon, I.-J., I. Ginis, T. Hara, H. L. Tolman, C. W. Wright, and E. J. Walsh, 2003: Numerical simulation of sea surface directional wave spectra under hurricane wind forcing. J. Phys. Oceanogr., 33, 1680–1706, https://doi.org/10.1175/2410.1.
Moon, I.-J., T. Hara, I. Ginis, S. E. Belcher, and H. L. Tolman, 2004: Effect of surface waves on air–sea momentum exchange. Part I: Effect of mature and growing seas. J. Atmos. Sci., 61, 2321–2333, https://doi.org/10.1175/1520-0469(2004)061<2321:EOSWOA>2.0.CO;2.
Moon, I.-J., J.-I. Kwon, J.-C. Lee, J.-S. Shim, S. K. Kang, I. S. Oh, and S. J. Kwon, 2009: Effect of the surface wind stress parameterization on the storm surge modeling. Ocean Modell., 29, 115–127, https://doi.org/10.1016/j.ocemod.2009.03.006.
Mueller, J. A., and F. Veron, 2009: Nonlinear formulation of the bulk surface stress over breaking waves: Feedback mechanisms from air-flow separation. Bound.-Layer Meteor., 130, 117–134, https://doi.org/10.1007/s10546-008-9334-6.
Peirson, W. L., and A. W. Garcia, 2008: On the wind-induced growth of slow water waves of finite steepness. J. Fluid Mech., 608, 243–274, https://doi.org/10.1017/S002211200800205X.
Peirson, W. L., A. W. Garcia, and S. E. Pells, 2003: Water wave attenuation due to opposing wind. J. Fluid Mech., 487, 345–365, https://doi.org/10.1017/S0022112003004750.
Powell, M. D., P. J. Vickery, and T. A. Reinhold, 2003: Reduced drag coefficient for high wind speeds in tropical cyclones. Nature, 422, 279–283, https://doi.org/10.1038/nature01481.
Reichl, B. G., T. Hara, and I. Ginis, 2014: Sea state dependence of the wind stress over the ocean under hurricane winds. J. Geophys. Res. Oceans, 119, 30–51, https://doi.org/10.1002/2013JC009289.
Reul, N., H. Branger, and J.-P. Giovanangeli, 2008: Air flow structure over short-gravity breaking water waves. Bound.-Layer Meteor., 126, 477–505, https://doi.org/10.1007/s10546-007-9240-3.
Richter, D. H., and P. P. Sullivan, 2013: Sea surface drag and the role of spray. Geophys. Res. Lett., 40, 656–660, https://doi.org/10.1002/grl.50163.
Roekel, L. V., B. Fox-Kemper, P. P. Sullivan, P. E. Hamlington, and S. Haney, 2012: The form and orientation of Langmuir cells for misaligned winds and waves. J. Geophys. Res., 117, C05001, https://doi.org/10.1029/2011JC007516.
Savelyev, I. B., M. P. Buckley, and B. K. Haus, 2020: The impact of nonbreaking waves on wind-driven ocean surface turbulence. J. Geophys. Res. Oceans, 125, e2019JC015573, https://doi.org/10.1029/2019JC015573.
Snyder, R. L., F. W. Dobson, J. A. Elliott, and R. B. Long, 1981: Array measurements of atmospheric pressure fluctuations above surface gravity waves. J. Fluid Mech., 102, 1–59, https://doi.org/10.1017/S0022112081002528.
Sullivan, P. P., and J. C. McWilliams, 2010: Dynamics of winds and currents coupled to surface waves. Annu. Rev. Fluid Mech., 42, 19–42, https://doi.org/10.1146/annurev-fluid-121108-145541.
Sullivan, P. P., J. C. McWilliams, and C.-H. Moeng, 2000: Simulation of turbulent flow over idealized water waves. J. Fluid Mech., 404, 47–85, https://doi.org/10.1017/S0022112099006965.
Sullivan, P. P., J. C. McWilliams, and E. G. Patton, 2014: Large-eddy simulation of marine atmospheric boundary layers above a spectrum of moving waves. J. Atmos. Sci., 71, 4001–4027, https://doi.org/10.1175/JAS-D-14-0095.1.
Sullivan, P. P., M. L. Banner, R. P. Morison, and W. L. Peirson, 2018a: Impacts of wave age on turbulent flow and drag of steep waves. Procedia IUTAM, 26, 174–183, https://doi.org/10.1016/j.piutam.2018.03.017.
Sullivan, P. P., M. L. Banner, R. P. Morison, and W. L. Peirson, 2018b: Turbulent flow over steep steady and unsteady waves under strong wind forcing. J. Phys. Oceanogr., 48, 3–27, https://doi.org/10.1175/JPO-D-17-0118.1.
Suzuki, N., Y. Toba, and S. Komori, 2010: Examination of drag coefficient with special reference to the wind sea Reynolds number: Conditions with counter and mixed swell. J. Oceanogr., 66, 731–739, https://doi.org/10.1007/s10872-010-0060-6.
Suzuki, N., T. Hara, and P. P. Sullivan, 2013: Impact of breaking wave form drag on near-surface turbulence and drag coefficient over young seas at high winds. J. Phys. Oceanogr., 43, 324–343, https://doi.org/10.1175/JPO-D-12-0127.1.
Teixeira, M., 2018: A model for the wind-driven current in the wavy oceanic surface layer: Apparent friction velocity reduction and roughness length enhancement. J. Phys. Oceanogr., 48, 2721–2736, https://doi.org/10.1175/JPO-D-18-0086.1.
Tolman, H. L., and D. Chalikov, 1996: Source terms in a third-generation wind wave model. J. Phys. Oceanogr., 26, 2497–2518, https://doi.org/10.1175/1520-0485(1996)026<2497:STIATG>2.0.CO;2.
Troitskaya, Y., D. Sergeev, O. Ermakova, and G. Balandina, 2011: Statistical parameters of the air turbulent boundary layer over steep water waves measured by the PIV technique. J. Phys. Oceanogr., 41, 1421–1454, https://doi.org/10.1175/2011JPO4392.1.
Troitskaya, Y., D. A. Sergeev, O. Druzhinin, A. A. Kandaurov, O. S. Ermakova, E. V. Ezhova, I. Esau, and S. Zilitinkevich, 2014: Atmospheric boundary layer over steep surface waves. Ocean Dyn., 64, 1153–1161, https://doi.org/10.1007/s10236-014-0743-4.
Veron, F., 2015: Ocean spray. Annu. Rev. Fluid Mech., 47, 507–538, https://doi.org/10.1146/annurev-fluid-010814-014651.
Veron, F., G. Saxena, and S. K. Misra, 2007: Measurements of the viscous tangential stress in the airflow above wind waves. Geophys. Res. Lett., 34, L19603, https://doi.org/10.1029/2007GL031242.
Wang, D., T. Kukulka, B. G. Reichl, T. Hara, and I. Ginis, 2019: Wind–wave misalignment effects on Langmuir turbulence in tropical cyclone conditions. J. Phys. Oceanogr., 49, 3109–3126, https://doi.org/10.1175/JPO-D-19-0093.1.
Yang, D., and L. Shen, 2010: Direct-simulation-based study of turbulent flow over various waving boundaries. J. Fluid Mech., 650, 131–180, https://doi.org/10.1017/S0022112009993557.
Yang, D., and L. Shen, 2017: Direct numerical simulation of scalar transport in turbulent flows over progressive surface waves. J. Fluid Mech., 819, 58–103, https://doi.org/10.1017/jfm.2017.164.
Yang, D., C. Meneveau, and L. Shen, 2013: Dynamic modelling of sea-surface roughness for large-eddy simulation of wind over ocean wave field. J. Fluid Mech., 726, 62–99, https://doi.org/10.1017/jfm.2013.215.
Yang, Z., B.-Q. Deng, and L. Shen, 2018: Direct numerical simulation of wind turbulence over breaking waves. J. Fluid Mech., 850, 120–155, https://doi.org/10.1017/jfm.2018.466.
Young, I. R., and R. J. Sobey, 1985: Measurements of the wind-wave energy flux in an opposing wind. J. Fluid Mech., 151, 427–442, https://doi.org/10.1017/S0022112085001033.
Yousefi, K., F. Veron, and M. P. Buckley, 2020: Momentum flux measurements in the airflow over wind-generated surface waves. J. Fluid Mech., 895, A15, https://doi.org/10.1017/jfm.2020.276.
Zdyrski, T., and F. Feddersen, 2020: Wind-induced changes to surface gravity wave shape in deep to intermediate water. J. Fluid Mech., 903, A31, https://doi.org/10.1017/jfm.2020.628.