<<

870 JOURNAL OF PHYSICAL VOLUME 44

Langmuir Turbulence in Swell

JAMES C. MCWILLIAMS,EDWARD HUCKLE, AND JUNHONG LIANG Department of Atmospheric and Oceanic Sciences, University of California, Los Angeles, Los Angeles, California

PETER P. SULLIVAN National Center for Atmospheric Research, Boulder, Colorado

(Manuscript received 29 May 2013, in final form 14 November 2013)

ABSTRACT

The problem is posed and solved for the oceanic surface boundary layer in the presence of wind stress, stable density stratification, equilibrium wind-waves, and remotely generated swell-waves. The addition of swell causes an amplification of the Lagrangian-mean and rotation toward the swell-wave direction, a fat- tening of the Ekman velocity spiral and associated vertical Reynolds stress profile, an amplification of the inertial current response, an enhancement of turbulent variance and entrainment rate from the pycnocline, and—for very large swell—an upscaling of the coherent patterns. Implica- tions are discussed for the parameterization of Langmuir turbulence influences on the mean current profile and the material entrainment rate in oceanic circulation models. In particular, even though the turbulent kinetic energy monotonically increases with wave amplitude inversely expressed by the turbulent Langmuir number La, the Lagrangian shear profile kL(z) is a nonmonotonic function of La, first increasing with increasing wave amplitude up to approximately the wind-wave equilibrium level, then decreasing with ad- 2 ditional swell-wave amplitude. In contrast, the pycnocline entrainment rate is a monotonic function ;La 2.

1. Introduction The physics of this regime is now well understood, primarily on the basis of many large eddy simulation The wind blows and the waves rise and roll on. This is (LES) studies of the wind-wave equilibrium state [e.g., the regime of Langmuir turbulence in the oceanic sur- the review in Sullivan and McWilliams (2010)]. How- face boundary layer (BL), so-called because Langmuir ever, in nature the wind is rarely in equilibrium with the circulations (often recognized by the windrows in the waves (Young 1999; Sullivan et al. 2008; Hanley et al. they cause) are the primary turbulent eddies 2010). Often this is due to transient wind changes or whose vertical momentum and buoyancy fluxes main- limited fetch, when the waves have not yet evolved into tain the mean ageostrophic current and density stratifi- a fully developed equilibrium but are headed there over cation. Langmuir circulations arise from the instability an interval of hours or over an offshore distance of of wind-driven boundary layer shear in the presence of tens of kilometers. Disequilibrium also occurs when (Craik and Leibovich 1976; McWilliams swell-waves with a large amplitude and long wavelength et al. 1997). Alternatively expressed, this regime is a propagate thousands of kilometers away from their wind-driven with stable interior density generation in an earlier strong storm to a place with stratification and surface gravity waves that induce weaker local wind whose wind-waves are weaker and wave-averaged vortex and and buoyancy shorter. During their propagation, swell waves exhibit advection due to the Lagrangian-mean Stokes drift slow decay and slow evolution through and velocity. Langmuir turbulence is common in nature four-wave resonant interaction, and they have weak (Belcher 2012). nonlinear interactions with the wind-waves encountered en route as long as there is significant separation in wavelength and/or direction (Masson 1993; Zakharov Corresponding author address: James C. McWilliams, Dept. of Atmospheric and Oceanic Sciences, UCLA, 405 Hilgard Ave., Los 2005; Depley et al. 2010). Under these conditions, a local Angeles, CA 90095-1565. quasi-equilibrium state of linearly superimposed swell- E-mail: [email protected] and wind-waves can persist for a day or more if the wind

DOI: 10.1175/JPO-D-13-0122.1

Ó 2014 American Meteorological Society MARCH 2014 M C W I L L I A M S E T A L . 871 is steady. These circumstances are especially common in friction velocity). Apart from the level of turbulent en- the tropics and subtropics in both hemispheres, with ergy and differences in the profiles of mean current and swell-waves coming from higher-latitude winter storms Reynolds stress very near the surface, this distinction (Alves 2006; Hanley et al. 2010). Also, they can occur in the representation of the breakers is not important anywhere in the lull after a storm. for the topics of this paper. In this paper, t and its as- Several idealizations of persistent disequilibrium have sociated hAi(z) are calculated from the near-surface 1 been examined with LES. Commonly, the wave spectrum wind speed Ua with a bulk drag law; that is, t 5 r 2 r 5 23 5 is simplified to a monochromatic wave at the spectrum aCDUa , with a 1kgm and CD 1.3 for the cases 21 peak (McWilliams et al. 1997), the wind-wave Stokes drift here with Ua # 10 m s . is multiplied by a factor (usually less than one, implying The LES model is spun up from rest and approxi- weaker waves with a turbulent Langmuir number La $ mately equilibrates over a period of a day or so, while 0.3, its equilibrium value; see section 2) (Harcourt and retaining inertial oscillations that slowly decay over D’Asaro 2008; Grant and Belcher 2009), or the wind- many days. In the present problem, an initial mean wave Stokes drift direction is rotated relative to wind stratification T(z) is included. It initially alignment (Van Roekel et al. 2012). All of these formu- has a well-mixed layer in the top 33 m and an interior 2 lations are ad hoc compared to realistic persistent wind gradient of 0.18Cm 1 (implying a buoyancy frequency 2 regimes where the wind- evolves toward equilibrium. of N 5 0.014 s 1). Turbulent entrainment develops, and On the other hand, the general formulation of wind-wave the boundary layer deepens. Because this N value is disequilibrium is daunting in the complexity of possible rather large, the boundary layer depth h(t) only slowly transient histories. In contrast, the case of equilibrium deepens while the average mixed layer temperature T(t) wind-waves with added steady swell-waves with well- cools in the absence of a surface heat flux. The Coriolis 2 2 separated spectrum peak wavenumbers (hence La , 0.3) frequency f is 10 4 s 1 (corresponding to a latitude of is an apt idealization of realistic persistent situations. 458N) and is assumed to be spatially uniform over the Because the remote swell-generating storm and the local small spatial scale of the domain. With the traditional wind are independent, the added swell component can approximation that neglects the horizontal projection have an arbitrary orientation and amplitude relative to of Earth’s rotation vector, the are symmetric the local wind-sea. This is the problem addressed here. with respect to wind direction; hence, without loss of generality, the wind direction is chosen as x^ [east (E)] and held steady. The wind-sea spectrum is in equilib- 2. Formulation rium with the wind according to the prescription of The LES code solves the wave-averaged dynamical Alves et al. (2003). The associated profiles for the w equations in Sullivan and McWilliams (2010). These Stokes drift velocity ust(z) and breaker forcings are w are incompressible, rotating Boussinesq fluid dynamics calculated as in Sullivan et al. (2007). The ust is aligned with a prognostic turbulent kinetic energy equation for with the wind direction. subgrid-scale fluxes. Both model components have sur- The preceding formulation yields familiar face wave effects that include Stokes drift Coriolis and behaviors that are described in several papers cited vortex forces and scalar advection, augmented pres- above. The novel aspect here is an added swell-wave sure head, and subgrid Stokes drift energy production component to the Stokes drift, which is modeled as and advection. The ‘‘wind’’ forcing is by parameterized h i z breakers that inject a body A and subgrid work us 5 Us exp u^s (1) W; for simplicity in the present paper, ensemble-mean st Ds breaker forcing profiles hAi(z) and hWi(z) are used, rather than stochastic individual impulses. (Angle brack- for z # 0 where mean is z 5 0. Because swell- ets indicate a horizontal and temporal average, and z is wave spectra are narrow, the swell is represented as the vertical coordinate.) The same dynamical formulation a linear, monochromatic surface with period is used in McWilliams et al. (2012) for the wavy Ekman P, wavelength l,andamplitudepffiffiffiffiffiffiffiffiffiffiffiffiffia. The dispersion rela- layer with uniform density and equilibrium wind-waves. tion in deep water is P 5 2pl/g (g is gravitational ac- In particular, in that paper the differences between en- celeration). The associated Stokes drift has a depth semble and stochastic breaker forcing are analyzed, as scale Ds 5 gP2/8p2 and a surface velocity Us 5 8p3a2/gP 3. well as their differences compared with a conventional surfaceÐ stress forcing with an equivalent integral effect h i 5 t r 5 2 1 (i.e., A dz / o u* in wind-wave equilibrium, where A more accurate drag law should include the influence of swell t r is stress, o is the mean density, and u* is the oceanic waves on CD for low wind speed (Sullivan et al. 2008). 872 JOURNAL OF VOLUME 44

FIG. 1. Wave spectra during mixed wind and swell-wave conditions at the Harvest buoy (near Point Conception, California) from CDIP at 0030 UTC 5 Dec 2007. (left) Azimuthally integrated frequency spectrum for sea surface elevation F(v), and (right) normalized directional spectrum D(v, f) 5 G(v, f)/F(v), where G is the directional 2 spectrum (note that its f integral is unity for each v). The v (rad s 1) is the wave frequency, and f is the wave 21 propagation direction. The simultaneous Ua is 12 m s in the direction of the black arrow, measured at the closest National Data Buoy Center station 46054. The wind-sea has a peak P of about 4.5 s to the SE, and the swell wave has

a P of 19 s to the E. The Hs is 6.9 m.

The term u^s is the unit direction vector for swell-wave swell-wave direction. Figure 1 shows a particular example propagation. The total Stokes drift is the sum of the wind- of mixed wind- and swell-waves with large Hs. sea and swell contributions assuming no nonlinear in- LES solutions are calculated in a vertical domain H 5 teraction between their spectral components; that is, 100 m thick and a horizontally periodic domain (of width L 5 300 m, except where otherwise stated). The vertical 5 w 1 s ust(z) ust(z) ust(z). (2) grid is stretched with a smallest spacing of 0.3 m near the surface. The horizontal grid is uniform with a spacing of Wave measurements are routinely made from oceanic 2 m. These grids encompass the boundary layer turbu- buoys, and an extensive compendium is available from lence down to the resolution limit where the subgrid the Coastal Data Information Program (CDIP; www. parameterization regularizes the flow. cdip.ucsd.edu). Data and plots are presented in terms of The influence of swell-waves is explored through its 5 5 v p s s u^s P (orpffiffiffif 1/P /2 ) and significant Hs parameters U , D , and (Table 1). In the parameter (5 2 2a for a monochromatic wave with sea level am- survey case sets, these are varied independently; how- plitude a). In some analyses waves with a spectrum peak ever, in nature Us and Ds tend to be anticorrelated be- P longer than 10 s are categorized as swell-waves, al- cause of their inverse dependencies on powers of the though they are wind-waves in the midst of high winds; period P [see just below (1)]. The LES results are non- this ambiguity is rarer at subtropical and tropical sites dimensionalized by u and by h diagnosed from the * with fewer big storms. Climatological histograms for solution averaged over the statistical analysis period t 5 individual sites are presented for the joint distribution of 0.55 2 1.18 3 105 s (i.e., one inertial period, 2p/f, with 24 21 P and Hs.EventswithHs . 10 m or P . 20 s are found, a Coriolis frequency f 5 10 s ). The term h(t) is de- although the bulk of the swell-wave events are not so fined as the depth at which the mean thermal gradient s 21 extreme. These imply possible ranges of U up to ›zhTi(z) becomes as large as 0.058Cm just below the 2 0.4 m s 1 and Ds up to 50 m, both of which are much well-mixed layer. Initially, h(0) 5 33 m in all cases. larger than the amplitude and depth scale for Stokes drift An important parameter for Langmuir turbulence is in a moderate wind-sea. No restriction is placed on the the so-called Langmuir number: MARCH 2014 M C W I L L I A M S E T A L . 873

w TABLE 1. Case sets analyzed in this paper. In all cases ust and As is commonly done for analytic convenience, this is (A, W) are in equilibrium with the eastward wind. For these case rewritten as a scalar equation for the complex velocity sets, the La values are 0.14–0.24 (Us1), 0.13–0.29 (Us2), 0.16 (Ds), U 5 hui 1 ihyi: 0.16–0.49 (Qs), 0.11–0.16 (Ua), and 0.11 (Lh). In addition, there is a comparison case NB (i.e., No Stokes drift velocity and ensemble- 5 21 5 5‘ U 1 U 5 k ›2U mean Breaker forcing) with Ua 10 m s and ust 0 (La ). if ( st) o z , (5)

Case set Wind and swell-wave parameters x y with boundary conditions of ko›zU(0) 5 T [ (t 1 it )/ 5 21 s5 us5 s 5 21 Us1 Ua 10 m s ; D 16 m; 0; U 0–0.4 m s r U 2‘ 5 21 s s s 21 o and ( ) 0. The problem is further simplified by Us2 Ua 5 5ms ; D 5 16 m; u 5 0; U 5 0–0.25 m s 21 s s assuming that the Stokes drift (2) has a monochromatic Ds Ua 5 10 m s ; D 5 4–24 m (P 5 6–14 s); u 5 0; 2 Us5 0.25 m s 1 wind-wave component (identified with the wind-wave 21 s s Qs Ua 5 10 m s ; D 5 16 m; u 52p 21p; spectrum peak), structurally analogous to the swell- 2 Us 5 0.25 m s 1 wave component (1), so that 21 s s Ua Ua 5 2.5–10 m s ; D 5 16 m; u 5 0; h i h i 2 z z Us 5 0.25 m s 1 U [ u 1 iy 5 Uw exp 1 Us exp[ius] exp . 21 s s s 21 st st st w s Lh Ua5 2.5 m s ; D 5 16 m; u 5 0; U 5 0.25 m s ; D D L 5 300–1200 m (6) To match the problem posed in section 2, T, Uw, and Us   1/2 are taken to be real and positive. Also, assume that f . 0 u* La 5 (3) (Northern Hemisphere). ju (0)j st Using simple analytic techniques, the Ekman–Stokes [sometimes called the turbulent La to distinguish it from solution is derived: the laminar La defined for viscously controlled flows;   h i (1 1 i)z ifUw z U 5 U exp 1 exp Leibovich (1983)]. Small La is associated with large e k w2 2 w he o/D if D waves, especially large swell waves. Here, the surface h i 5 ifUs exp[ius] z value of ust is evaluated at the first grid level of z 1 exp , (7) 2 k s2 2 s 0.15 m. The value of La is weakly grid dependent be- o/D if D cause there is a logarithmic singularity in uw,asz / 0 pffiffiffiffiffiffiffiffiffiffiffi st 5 k arising from the high-frequency tail in the wind-sea with Ekman depth he 2 o/f and amplitude coefficient spectrum of Alves et al. (2003) with F(f ) } f 24. However, ( ) w s s Romero and Melville (2010) show that the tail steepens 12 i ik fU ik fU exp[iu ] U 5 pffiffiffiffiffiffiffiffiffiffi T 2 o 2 o . } 25 e k w k w2 2 s k s2 2 to F f at sufficiently large f because of saturation, 2f o D ( o/D if ) D ( o/D if ) w thus regularizing ust (0) on a vertical scale finer than re- (8) s / solved here. The swell-wave ust in (1) is smooth as z 0. Notice that a vector misalignment between swell- and The transport integrals are wind-waves causes a reduction in ju (0)j, hence an in- st ð crease in La, compared to alignment. Depending on the 0 T i w w s s s swell amplitude and orientation, La may or may not be U dz 52 2 D U 2 D U exp[iu ], 2‘ f increased compared to the case of only wind-waves. ð 0 U dz 5 DwUw 1 DsUs exp[ius], and 2‘ st 3. Analytic Ekman–Stokes model ð 0 iT UL dz 52 , (9) A partial interpretation of the LES results in sections 4 2‘ f and 5 comes from an analytic solution of a wave-modified L extension of the classic Ekman layer model, which is now where the complex Lagrangian velocity is U 5 U 1 Ust L called the Ekman–Stokes model. It represents a horizontal (i. e. , u 5 hui 1 ust in vector notation). The Lagrangian and time average of the horizontal momentum balance transport is rotated 908 to the right of the wind stress, while (14) with three simplifying assumptions: constant density the Eulerian transport also has anti-Stokes components. s w ro, constant eddy viscosity ko, and surface wind stress t For the usual large swell situation of U . U and instead of breaker impulse A: Ds . Dw, these solutions indicate that a sufficient con- dition for the local wind-wave influence to be relatively f^z 3 (hui 1 u ) 5 k ›2hui, (4) st o z unimportant in the boundary layer flow profile is for k › h i 5 t r 5 s s w w s 6¼ with a surface boundary condition of o z u (0) / o U D U D . All of the cases in Table 1 with U 0 are 2 / / 2‘ u* and a lower boundary condition of u 0asz . in this regime. 874 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 44

FIG. 2. (left) Eastward and (right) northward velocities for two cases in the case set Us1: wind-wave equilibrium 2 (black) and with an added swell with Us 5 0.25 m s 1, Ds 5 16 m, and a wind-aligned us 5 0 (red). The solid lines are

the mean velocity, and the dashed lines are the Stokes drift ust.

The solutions (7) and (8) have a left–right symmetry inertial current amplitude is made with the steady velocity with respect to the swell direction us relative to the wind averaged over the boundary layer, that is, the transport direction (here E). Define U^ as the solution without magnitude divided by the Ekman depth scale s 5 s 6¼ us swell (i.e., for U 0). Then for U 0 and general , ð 0 V s 5 U s 2 U^ the difference fields u u are equivalent when 5 1 U us Uin 0 dz , (12) rotated by the horizontal angle 2 ; that is, he 2‘ V 5 us V us exp[i2 ] 2us . (10) using (9) to evaluate the integral. For comparison with the LES results below, the 2k › U The same rotational symmetry holds for o z ,which Ekman–Stokes solution is normalized by u* for velocity, is the Ekman–Stokes equivalent of Reynolds stress minus 2 he for depth, and u* for the equivalent of Reynolds stress the upward-integrated breaker forcing [i.e., R(z) in (15)]. (i.e., 2ko›zU or 2ko›zu). As will be shown, the matches For flows impulsively started from rest, the preceding of the Ekman–Stokes solutions to the LES results are not steady solution U(z) is augmented by an inertial oscil- precise in their vertical profile structure, as is to be ex- lation Uin(z, t) such that their sum is zero at t 5 0 at all pected with the strong simplifications of uniform density depths. With an additional acceleration term ›t U in (5), and eddy viscosity (cf. Fig. 6, described in greater detail w the homogeneous solutions are the eigenmodes below) and a monochromatic wind-sea profile for ust(z) (cf. Fig. 2). The Ekman–Stokes parameters are chosen, U } 2 2 k 2 $ first, by exactly matching u , Us, us,andDs with their LES in exp[ ift ok t] cos[kz], k 0, (11) * values. Second, the measured LES boundary layer depth which imply vertical oscillation with wavenumber k, ph ffiffiffiffiffiffiffiffiffiffiffi(which is limited by stable stratification rather than anticyclonic temporal rotation of the current direction at 2ko/f for he) is used for its normalizations of z and an inertial frequency f, and amplitude decay at a rate Reynolds stress. Third, for the remaining Ekman–Stokes 2 w w kok . The longer-lasting components have smaller k, and parameters—U , D ,andko—the simple profile assump- hence a larger vertical scale. So an estimate of the initial tions for the wind-sea Stokes drift velocity and diagnosed MARCH 2014 M C W I L L I A M S E T A L . 875

L FIG. 3. Hodographs of Lagrangian-mean velocity, u (z) in (13), for the case sets (left) Us1 and (right) Qs. The dots along the spirals indicate depths of 2z/h 5 (0.1, 0.3, 0.6, 0.9). The arrow indicates increasing La from the largest (outer spiral) to no swell (inner spiral) (left). The us values are labeled by their compass directions, and, for clarity, the us $ 0 are separated from the us # 0 cases (right). The insets show the equivalent spirals from the Ekman–Stokes model, with the us 56p (W) profile having very small uL(z) values at all depths. eddy viscosity do not correspond to the more complex dependencies exhibited in the more complete LES solu- profiles in the LES solution. Therefore, the Ekman– tions, and by this standard it proves to be useful. 2 Stokes parameter values are chosen—0.125 m s 1,1.25m, 2 and 0.025 m2 s 1, respectively—to give the best visual 4. Mean current and Reynolds stress comparisons below in Figs. 3–5. These parameter values w k are not ill matched to the more complex ust(z)and (z) From a larger-scale perspective, the most important profiles in Figs. 2 and 6, respectively. Furthermore, the outcomes for the boundary layer are the quasi-steady normalized Ekman–Stokes profile shapes are not highly mean horizontal current and slowly eroding density strat- sensitive to these parameter choices. At best, the Ekman– ification after an initial spinup period of about a day. After s Stokes model is a simple explanation for some of the ust filtering out the inertial current that arises from the

FIG. 4. Hodographs of the Reynolds stress plus breaker impulse, R(z) in (15), for the case sets (left) Us1 and (right) Qs. The format is the same as in Fig. 3. For the insets, the comparable quantity in the Ekman–Stokes model is 2ko›su(z): 52 2 / R(0) u* at the surface, and R 0 at depth. 876 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 44

FIG. 5. Inertial current amplitude Uin 0 averaged over the boundary layer (2h # z # 0) as a function of (top, left) La for case sets Us1 (black), Us2 (red), and Ua (blue); as a function of (top, right) us for case set Qs; and as a function of (bottom) Ds for case set Ds. The insets show the equivalent results from the Ekman–Stokes model using (12). impulsive start to the boundary layer flow, the mean cur- a bit larger through enhanced entrainment (section 6). rent profile hui(z) in Fig. 2 is a combination of a clockwise- Although such large swell has not been examined be- rotating in uL and an anti-Stokes Eulerian fore, these behaviors are not surprising for Langmuir

flow with hui opposing ust, as in the Ekman–Stokes solu- turbulence solutions. tion (7). Both features are familiar attributes of equilibrium The anti-Stokes effect is partly canceled in the Langmuir turbulence (e.g., McWilliams et al. 1997); for Lagrangian-mean flow profile example, if u 5 0 (as in case NB defined in Table 1; not st uL(z) 5 hui 1 u . (13) shown), the Eulerian hui . 0 near the surface because st there is no anti-Stokes effect. With an added swell aligned In the Ekman–Stokes model the cancellation is com- with the wind, the Stokes drift velocity is stronger and plete in the transport (9), as is also true in the LES that deeper. The wind-aligned mean velocity is increasingly satisfies the same transport relation (16). When the swell upwind, the crosswind velocity is more nearly linear in its amplitude Us increases for wind-aligned swell (i.e., us 5 0; profile and reverses near the boundary layer base, and Fig. 3, left), the surface maximum in uL is larger, and both components extend somewhat deeper because h is the Ekman spiral is fatter; that is, the flow in the bulk of MARCH 2014 M C W I L L I A M S E T A L . 877

L L L FIG. 6. Hodographs of the eddy viscosity vector, k 5 (kk , k?) in (18) normalized by u h, for the case sets (left) Us1, (right) Qs, and * 2 (bottom) Ds. The format is similar to Fig. 3. (top, left) The cases are for Us 5 0 (blue), 0.1 (red), 0.25 (black), and 0.4 m s 1 (green), plus the case NB with ust 5 0 (defined in Table 1; black dashed); the curved arrow indicates the direction of increasing La. The loops are traversed counterclockwise with depth from the origin out to a middepth max and back to the origin. The Ekman–Stokes model (section 3) would be represented as a point on the abscissa at (k , 0)/u h , and the standard K-profile parameterization (KPP) scheme (Large et al. 1994) would o * e be a line traversed by (K(z), 0)/u h, with K(z) a convex positive function of z that starts and ends near the origin. * the boundary layer is stronger. Compared to case NB parameter values (section 3), which is plotted in the in- without Stokes drift (defined in Table 1; not shown), set. When the swell direction us varies away from wind these wave-influenced cases have a stronger surface uL alignment to the east (Fig. 3, right), uL at the surface and a fatter Ekman. The same qualitative trend with La weakens and rotates toward the swell, but only over a occurs in the Ekman–Stokes solution for analogous limited azimuthal range. In the interior of the boundary 878 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 44 layer, the upwind velocity component is strongest—the hodograph becomes quite thin. These dependencies on Us Ekman spiral is fattest—in the us range from southeast and us are also present in the Ekman–Stokes model (in- (SE) to northeast (NE) and is much weaker over the rest sets), although the shapes of its hodographs are not the of the compass range. Again, these behaviors are similar same, and the left–right symmetry in us is not as visually to those in the Ekman–Stokes model (shown as insets in apparent, in spite of the fact that it does have the partic- Figs. 3, 4, and 5). In Fig. 3, there is a rough left–right ular us symmetry (10). Again, the Ds dependencies of the symmetry with a sign reversal in us [e.g., north (N) versus Reynolds stress (not shown) are much smaller than those south (S)]. The dependency of uL(z) on the swell depth for Us and us. scale Ds (i.e., in the case set Ds) is rather slight and is not The mean momentum balance (14) may equivalently shown; with a smaller Ds, uL(z) is slightly more surface be expressed in terms of a Lagrangian eddy viscosity intensified. Of course, the anti-Stokes aspect in hui is tensor defined by larger and reaches deeper with larger Ds, as in (7). ! The mean horizontal momentum balance (averaged cosuL 2sinuL 2h 0 0i 5 kLR › L R 5 u w zu , , (17) over the inertial oscillation) is sinuL cosuL ^ 3 h i 1 52› h 0 0i 1 ^ f z ( u ust) z u w xA, (14) where R(u) is the horizontal rotation matrix represent- ing the rotation of the shear direction into the opposite 0 0 with a surface boundary condition of hu w i(0) 5 0 due of the Reynolds stress direction. Here, kL(z) is the to the replacement of wind stress by the mean breaker positive scalar magnitude, and uL(z) is the rotation an- momentum impulse A(z) . 0 in a thin layer near the gle. A conventional eddy viscosity model has uL [ 0, but surface. The velocity on the left side is the Lagrangian- this is not sufficient to characterize an Ekman layer with L mean flow u (z), and one sees the constraint that an surface waves (McWilliams et al. 2012, section 3d). Al- anti-Stokes Eulerian flow hui develops with an in- ternatively, an eddy viscosity vector is defined in terms creasing Stokes drift ust and fixed momentum forcing of components of the Reynolds stress parallel and per- by the wind, with the Reynoldsstressprofileunlikely pendicular to the mean Lagrangian shear: to exceed the surface stress. The right side forcing in (14) is minus the divergence of the difference between › uL hu0w0i kL 5 kL uL 52 z the Reynolds stress and the upward integral of the breaker k (z) cos[ ] 2 , (› uL) impulse: z › L 3 h 0 0i ð L L L zu u w z k?(z) 5 k sin[u ] 52^z . (18) 5 h 0 0i 2 ^ 0 › L 2 R(z) u w x Adz . (15) ( zu ) 2H In the conventional model, kL . 0 and kL 5 0. Because of the boundary condition on Reynolds stress k ? In McWilliams et al. (2012) it is shown that a and the normalization for A, this has a surface value of Lagrangian eddy viscosity representation has simpler R(0) 52u2 . Thus, the transport integral, * vertical structure in k(z) and a smaller range of u(z) ð ð 0 0 2 variation than a conventional Eulerian one (i.e., defined 1 0 0 u uL dz 52 ^z 3 x^ A(z ) dz 52y^ * , (16) with uL replaced by hui in the preceding formulas) for 2 f 2 f H H unstratified, wind-wave equilibrium Langmuir turbu- lence, especially in the near-surface region where u is is independent of the swell amplitude [cf. (9)], despite st not small. The same is true in the present stratified cases its evident influence on the profile shape of uL(z) with swell-waves.2 The hodograph of the eddy viscosity (Fig. 3). vector kL(z) 5 (kL(z), kL(z)) (where the parentheses are The Reynolds stress hodograph for R(z) in Fig. 4 be- k ? standard math notation for expressing components of comes much fatter when Us is large and us is in the same a vector in this type of serial form) in Fig. 6 (left) shows quadrant as the wind direction. With large enough Us both the increase in the magnitude of kL in a wind-sea (which occurs in the case set Us2, not shown), the mag- compared to the absence of waves (case NB; defined in nitude of the crosswind hy0w0i even exceeds the sur- Table 1) and the ensuing decrease with added swell, face wind stress. For swell rotated to the left of the wind (us . 0), the R(z) hodograph tilts to the left and vice versa us , for 0; but again, the range in directional tilt is much 2 Similarly, Harcourt (2013) proposes a local second-moment us less than the range in . For the quadrants away from turbulence closure model based on Lagrangian-mean shear rather the wind [from south to north through west (W)], the than Eulerian. MARCH 2014 M C W I L L I A M S E T A L . 879

Us 6¼ 0. This nonmonotonic dependence on La is some- compared to a shear boundary layer without them (e.g., what surprising, and this is further discussed at the end of case NB; defined in Table 1). This is a different trend this section. The vertical profile of kL(z) has a middepth than the present result that kL decreases with the ad- maximum and vanishes at the edges of the boundary layer dition of an increasing swell-wave amplitude to an in all cases (i.e., the looping curves in the hodograph that equilibrium wind-wave field. The viscosity magnitude start and end at the origin). Without waves the loop is from (17) is the ratio of Reynolds stress and Lagrangian- very thin, indicating that uL ’ 0, but it is not thin for ei- mean shear magnitudes ther a wind-sea alone or a wind-sea with added swell. The 0 0 s 5 uL , jhu w ij wind-sea case with U 0has k 0 in the Stokes layer kL 5 , (19) uL . j› Lj near the surface and k 0 deeper in the interior of the zu boundary layer; the same structure occurs in the un- stratified, wavy Ekman layer (McWilliams et al. 2012). while the mean horizontal momentum balance (14) be- With increasing Us, the loops shrink and rotate clockwise low the wave-breaking layer implies in uL,sothatuL(z) , 0 at almost all depths with the s s largest U . With variable u (Fig. 6, right), the magnitude j› Lj ’ 1 j›2h 0 0ij u z u w . (20) of kL increases with leftward swell rotation up to us 5 p/2 z f (north), then begins to shrink again and uL rotates to the left. With rightward rotation in us, the magnitude de- Thus, us 52p kL creases to a minimum near /2 (south), and 0 0 s jhu w ij assumes a distended loop shape for u 523p/2 [south- kL ’ f . (21) s j›2h 0 0ij west (SW)]. This is a very substantial left–right u asym- z u w metry in the boundary layer response to swell. kL The Ekman–Stokes model (section 3) is not apt for To explain the trend in with decreasing La, for ex- ample, for wind-aligned waves, one can examine the these eddy viscosity behaviors because (ko, 0) is just a point on the positive abscissa of these hodographs. trends in the numerators and denominators of (19) or jh 0 0ij jh› Lij j›2h 0 0ij This implies that the analytic model is not reliable for (21). Both u w and zu , hence z u w as well, kL the detailed profile structure of the mean flow and are decreasing functions of depth, while their ratio Reynolds stress, even if it is useful for indicating their has a convex shape with a maximum near the middle of qualitative dependencies on the swell-wave parameters. the boundary layer (Fig. 6). With decreasing La, the h 0 0i In contrast, a ‘‘Lagrangian ’’ Ekman–Stokes u w (z) profile shows a systematically increasing mid- jh› Lij model—analogous to (5) and its boundary conditions depth numerator. The denominator zu first de- creases with the presence of weak waves, then around but with the Eulerian diffusive flux ko›z U replaced by L L ’ 3 a Lagrangian flux k ›zU —would be isomorphic to La 1 begins to increase with stronger waves, thus o kL a standard Ekman model without waves; that is, it would implying the reversing trends in . Consistent with (20), show no dependencies on the added swell and thus have the Reynolds stress profile curvature (i.e., its second no explanatory value for Figs. 3 and 4, as well as being derivative with depth) first diminishes with decreasing kL La, then because of the tendency of the hodograph to only a point located at ( o , 0) in Fig. 6. Finally, there is an appreciable dependency of kL(z) fatten (Fig. 4) with added swell, it also increases. The s k on D in the case set Ds (Fig. 6, bottom), with a larger Ekman–Stokes model with its fixed Eulerian o,in eddy viscosity magnitude as Ds is decreased, while main- contrast, has monotonically increasing numerators and taining a roughly similar loop shape as in Fig. 6 (left). denominators, and their middepth ratios as La decreases The complexity of the kL(z) profiles and their varia- (not shown). The implication is that, in the LES solution, tions with swell-wave parameters in Fig. 6 are chal- changes in the wave field change both the shape of lenging for achieving an accurate boundary layer the Langmuir circulations (section 7) with their signifi- parameterization for the shape of hui(z). On the other cant contribution to the momentum flux profile (e.g., hand, there is evident skill shown in the preceding demonstrated in McWilliams et al. 2012) and the out- figure insets for the very simple parameterization used in come of the anti-Stokes competition in the Lagrangian the Ekman–Stokes model with a constant ko (section 3). The implications of this trade-off between accuracy and 3 simplicity are further discussed in section 8. This result is based on LES solutions with an equilibrium wind- wave Stokes profiles whose amplitude is artificially reduced to in- It is known from previous work (McWilliams et al. crease La above its equilibrium value of 0.3. These solutions are not 1997; McWilliams and Sullivan 2000) that the eddy vis- shown here explicitly, but both Harcourt and D’Asaro (2008) and cosity magnitude increases due to the presence of waves Grant and Belcher (2009) show results for this larger La regime. 880 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 44 uL(z) profile in seemingly subtle ways that lead to the 6. Turbulent intensity, entrainment rate, diagnosed kL(z) changes with La. A more simplistic and energy production argument is that the numerator is plausibly con- 2 Define the turbulent kinetic energy profile as strained by the wind stress magnitude u*, while the denominators are freer to grow with increasing ›zust; 1 0 however, this argument does not faithfully capture the e(z) 5 (hu 2i 1 hw2i). (23) L decreasing j›zu j in the larger La regime. 2 The prime indicates the subtraction of the horizontally 5. Inertial current averaged flow at each time, that is, excluding both the As mentioned, the horizontally averaged current ex- mean and inertial currents. The e is the isotropic mea- hibits inertial currents generated in the initialization sure of turbulent intensity, and the vertical velocity shock. After adjustment, their vertical structure is ap- variance hw2i is useful to compare to float measure- proximately uniform in z over the boundary layer and ments in the surface boundary layer (Harcourt and weak below. For each LES case, a depth-averaged in- D’Asaro 2008). These two profile quantities are plotted ertial amplitude Uin 0 is fit during the analysis period in Fig. 7. They show a surface-intensified e(z)and of one inertial period. The dependencies with Us and us a shallow maximum in hw2i(z). With wind-aligned swell, are shown in Fig. 5. The inertial amplitude increases both e and hw2i increase compared to the wind-sea case; with Us (decreasing La), and the three case sets with this is similar to their increase in equilibrium Langmuir variable La (Us1, Us2, and Ua) all show a similar turbulence compared to a boundary layer without the functional shape. The us variations have an inertial Stokes vortex force (case NB defined in Table 1; not current amplitude dependency with a roughly a 1 b sin shown). Profiles are also shown for us 56p/2. For [us] dependency. These behaviors are qualitatively simi- northward swell, e is as large as with wind alignment, lar in the Ekman–Stokes (12), as is evident by comparison whereas for southward swell e is much reduced. Both 2 withtheinsetsinFig.5.TheUin 0 also increases approxi- e(z)andhw i(z) show some oscillatory complexity in mately linearly with Ds in the LES (Fig. 5, bottom), the northward case, including elevated variances in the consistent with the prediction in (12); that is, deep interior below the boundary layer (i.e., z ,2h) that indicate increased internal gravity wave excitation s s U 1 iT D U in 0 5 2 DwUw 2DsUs exp[ius] ’ (22) by the boundary layer turbulence impinging on the top u* u*h f u*h of the pycnocline (Polton et al. 2008). This is another asymmetric response to the swell angle for large Ds and Us. This relation explains the increase in us, and it is in the same sense as the stronger inertial s s s Uin 0 with decreasing La; that is, as D and U increase, La current response for u . 0 (section 5). However, there is decreases. It also explains the dependency on us.Whenus no temporal correlation of turbulent intensity with the is in the NE quadrant, with the minus sign in (22) the phase of the inertial current. This implies that the en- magnitude of Uin 0 increases in the first relation in (22) by hanced vertical shear at the base of the boundary layer the swell contribution being in between the southward when the inertial current aligns with the mean current contribution from wind stress and the westward negative shear is not an important source of turbulent energy Stokes drift contribution from the wind-sea. Conversely, production [in contrast to the situation of ‘‘inertial res- s when u is in the SW quadrant, the swell effect subtracts onance’’ when the wind rotates in phase with the inertial from the contributions from stress and wind-sea, making current (Price 1981)]. Uin 0 smaller. Overall, large swell amplifies the inertial To demonstrate the swell-wave parameter depen- 4 response to a sudden wind change. Furthermore, com- dencies, Fig. 8 shows e and hw2i after averaging both pared to case NB without Stokes drift (defined in Table 1; within the boundary layer and in the interior layer be- not shown), these wave-influenced cases have a stronger low. The turbulent intensities increase with smaller inertial current response. La in both layers, with the one exception of boundary layer hw2i for the most extreme case of weak winds in case set Ua. Through an argument that relates tur- 4 In LES cases that go beyond the largest Ds value in case set Ds, bulent intensity to the efficacy of Stokes production the further increases in Ds lead to little change in the turbulent 0 0 (i.e., Pst 52hu w i›zust . 0) in the turbulent kinetic intensity and fluxes. However, the anti-Stokes component in hui(z) continues to deepen proportionally, and the amplitude of the in- energy balance (Harcourt and D’Asaro 2008; Grant h 2i ; ertial response increases. No evident local instability below the and Belcher 2009), the scaling prediction is e, w 2 boundary layer arises in association with the deep hui(z) shear. La 4/3. This comparison curve is also drawn, and it is MARCH 2014 M C W I L L I A M S E T A L . 881

s FIG. 7. Turbulent variance profiles: (left) e(z) and (right) hw2i(z) for the case sets Us1 with U 5 0 (black) and Qs 2 with Us 5 0.25 m s 1 and us oriented toward east (red), north (green), and south (blue). roughly consistent with the La dependencies in the The swell-wave direction dependencies show the boundary layer for these cases with added swell that greatest inertial current and turbulent intensity for us in extend to smaller La than previously explored. The the NE quadrant, consistent with Figs. 5 and 7. The us increases in internal-wave variance below the bound- dependency in Fig. 8 is a roughly symmetric decrease ary layer are an even steeper function of inverse La, away from wind alignment for hw2i averaged over the 2 roughly ;La 2. boundary layer, although the vertical profile comparisons

FIG. 8. Depth-averaged turbulent variances: e (solid) and hw2i (dashed) grouped separately for within the BL (the upper curves; 2h # z # 0) and for the deep layer below (the lower curves; 2H # z # 2h). (left) A function of La is for the case sets Us1 (black), Us2 (red), and 2 Ua (blue). For comparison, a scaling curve ;La 4/3 is also plotted (light black; see text). (right) A function of us is for the case set Qs with the BL averages in blue and the deep averages in red. 882 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 44

FIG. 9. Entrainment-layer temperature flux extremum in z, normalized by 2u3 /agh (with thermal expansion coefficient a 5 2 3 2 2 * 10 4 8C 1), as a function of (left) La for case sets Us1 (black), Us2 (red), and Ua (blue) and as a function of (right) us for case set Qs. 2 A comparison curve ;La 2 (light gray) is drawn through the data end points (left).

s for u 56p/2 in Fig. 7 show that this symmetry is not agh2 agh 5 › PE [2 › T 52 min [hwTi(z)], (24) precise. The boundary layer e measures in Figs. 7 and 8 t 2 t 2 z are even farther away from us symmetry than hw2i,with us . larger e for 0. Below the boundary layer, both e and and then normalized by the scaling estimate for the rate 2 s hw i are so asymmetric in u that their maxima occur for 3 of wind work acting to increase the kinetic energy u*; northward swell (similar to the inertial current in Fig. 5). h i s that is, wT is normalized by a combined factor of The D dependencies in case set Ds (not shown) are 3 a a 2u*/ gh, with as the thermal expansion coefficient. rather slight, with some evident profile sensitivities in e(z) Figure 9 shows that added swell further enhances the and hw2i(z)andaweaktendencyfore averaged over the s entrainment rate, which increases as La decreases with boundary layer to increase as D decreases. a dependency that can be fit approximately as a linear 2 Langmuir turbulence is known to increase the en- function of La 2. The entrainment rate is largest trainment rate in a stratification-limited boundary layer when the swell is aligned with the wind-sea, with a sim- (McWilliams et al. 1997) as a consequence of Langmuir ilar us dependency as found for turbulent intensity e (i.e., circulations penetrating into the stable stratification and larger for us . 0 than for us , 0; Fig. 8). Consistent with scooping colder water into the boundary layer. This is the e change with Ds remarked on above, there is a weak the process that causes h(t) to increase in the present tendency for the minimum in hwTi to decrease in mag- solutions. nitude as Ds decreases. Both e and hw2i are larger with The entrainment rate is associated with the negative Stokes drift than without it (case NB defined in Table 1; minimum in the vertical temperature flux profile of not shown). Thus, stronger Langmuir turbulence has h i wT (z). This temperature flux extremum is rescaled by stronger entrainment. its contribution to the change in potential energy PE As expected, there is a strong correlation between associated with mixed layer deepening into a uniformly entrainment rate and the increased value of h averaged stratified interior and cooling of its average temperature over the analysis period. Because of the strong ther- T(t); that is, mocline, the changes in h are modest, ranging from only a slight increase over the initial h(0) 5 33 m in case NB

5 up to a value of 36 m for the wind-aligned case in case set Van Roekel et al. (2012, their Fig. 8) shows a monotonic de- s crease in hw2i(z) as a function of the Langmuir cell orientation Us1 with the strongest wind and swell. Yet again, the D angle, which is ’ us/2 for their misaligned cases. See section 7 for dependencies for hwTi and h(t) are modest, although further comparative remarks. there are small increases with Ds. MARCH 2014 M C W I L L I A M S E T A L . 883

LC FIG. 10. Vertical profiles of the orientation angle of Langmuir circulations (left) u (z) and their ellipticity (right) LC « (z) (ratio of major to minor axis) determined from fitting an ellipse to the horizontal correlation function Cw for w(x, y) at a fixed z in (27). These are for the cases from the case sets Us1 with Us 5 0 (black) and Qs with Us 5 2 0.25 m s 1 and us oriented toward east (red), north (green), and south (blue).

How can the us asymmetry in Figs. 8–10 be explained? rate for its linear shear instability that generates the 6 The ust profiles are symmetric with respect to the sign turbulence. Support for this conjecture is provided in of us, as part of the symmetry of the hui profiles in the Table 2, which has the normalized and depth-integrated Ekman–Stokes model (section 3). The answer must lie in values for the rates of mean shear and Stokes turbulent an asymmetry in the turbulent dynamics (cf. the constant kinetic energy production [(7) in McWilliams et al. (2012)]: ko assumption in Ekman–Stokes). A plausible cause is the difference in Eulerian-mean velocity hui(z)between us positive and negative . They are composed of an Ekman 6 To be specific we have in mind a calculation for given 1D spiral in the southern quadrants to the right of the wind profiles of hui(z), ustz, and hTi(z), with or without background plus an anti-Stokes component opposite to ust(z). For eddy viscosity and diffusivity profiles. These would be specified northward swell (us . 0), the southward anti-Stokes from the mean state of LES solutions of the type presented here. In component broadly reinforces the Ekman flow, whereas the case of nonrotating, unstratified flow, Craik and Leibovich (1976) and subsequent papers show that the Stokes vortex force for southward swell, the anti-Stokes component is op- enables a new shear instability mode as a paradigm for the emer- posing. Therefore, hui is stronger for northward swell gence of Langmuir circulations. Such a calculation has not yet been waves. We conjecture that this implies a greater growth performed for the conditions relevant here. 884 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 44

TABLE 2. Turbulent kinetic energy production rates of (25), ›zust, which is not as directly reduced by changes in the 3 us depth-integrated and normalized by u*. They are grouped by boundary layer turbulence, but would be so only if the Q magnitude for the case set s. feedback on the wave field, here neglected, were strong

Case Pu Pst Pu 1 Pst enough. This is a partial explanation for the strong inverse Us1: Us 5 0 (no swell) 1.5 13.2 14.7 dependence of e on La in Langmuir turbulence. Qs: us 5 0 (E) 3.2 21.3 24.5 Qs: us 5 p/4 (NE) 5.6 16.7 22.3 Qs: us 52p/4 (SE) 1.3 19.5 20.8 7. Langmuir circulations Q us 5 p s: /2 (N) 6.1 9.5 15.6 Langmuir circulations in a nonrotating, unstratified Qs: us 52p/2 (S) 0.8 12.9 13.7 Qs: us 5 3p/4 (NW) 8.5 3.3 11.8 surface layer have an idealized shape of closely packed, Qs: us 523p/4 (SW) 2.7 7.5 10.2 longitudinal roll cells, orienting with the surface stress Qs: us 56p (W) 7.7 3.1 10.8 and wind-sea and extending vertically throughout the boundary layer (Leibovich 1983). In equilibrium Langmuir 0 0 turbulence (McWilliams et al. 1997), they have more P (z) 52h w i› h i u u z u fragmented shapes but still retain appreciable horizontal P 52h 0 0i› anisotropy; they are rotated to the right of the wind-sea st(z) u w zust . (25) increasingly with depth. At sea they are most easily vi-

The mean shear production rate Pu, in particular, is sualized by buoyant debris trapped in surface conver- much larger for northward swell for a given value of jusj, gence lines, and in LES solutions the horizontal pattern of so that the total production rate is also larger. In almost w (especially w , 0) is a comparably useful visualization. all cases, Pu , Pst, with westward swell as the only ex- Statistical measures of the intensity and entrainment ception (where total production is small). Thus, stronger efficacy of Langmuir circulations show strong depen- hui is associated with a stronger turbulent production dencies on swell-wave amplitude and orientation (sec- rate, stronger turbulent intensity, and stronger entrain- tion 6). How does their spatial structure change with ment, and even with a fatter Reynolds stress profile swell? To analyze the horizontal structure of the 7 (Fig. 4), albeit only to a modest degree. Langmuir circulations, a spatial correlation function The total kinetic energy production (25) can alterna- is calculated from w(x, y) at each height z: tively be written using (17) and (19)–(21) as hw(x 1 j, y 1 h, z, t)w(x, y, z, t)i C (j, h, z) 5 , (27) P [ P 1 P 52h 0 0i› h Li w 2 u st u w z u hw(x, y, z, t) i 1 0 0 0 0 5 kLj› uLj cos[uL] 5 jhu w ij j›2hu w ij cos[uL]. (26) z f z where again the angle brackets denote an average over all (x, y, and t). Langmuir circulations have horizontally L Thus, the energy production is positive if cos[u ] . 0, as elongated patterns, which are expressed in Cw as ap- it is at all depths in Fig. 6, and it is proportional to the proximately elliptical contours decreasing away from product of the Reynolds stress and its profile curvature. the central extremum where Cw (0, 0, z) 5 1. Therefore, As discussed at the end of section 4, both of these factors an ellipse is fit to an intermediate contour of Cw in the 8 are increasing as La decreases in the swell regime (i.e., (j, h) plane (e.g., Cw 5 0.4 for Fig. 10). The direction of La # 0.3). This is consistent with energy production and its major axis is designated as uLC(z) and the ratio of velocity variance strongly increasing (Fig. 8), even while itsmajorandminoraxesas«LC(z). Respectively, these their ratio kL in (21) is decreasing (Fig. 6). Even for quantities measure the mean orientation angle and degree L weak waves with La . 1, the decrease in j›zu j is more of elongation (anisotropy) of the Langmuir circulations. than countered by the increase in jhu0w0ij (i.e., kL in- The resulting profiles of these measures are shown in creases), so that P also increases with decreasing La in Fig. 10 for several cases in case sets Us1 and Qs. For this regime. The increasing P is the plausible cause for the increasing e and entrainment as La decreases across its entire range. 7 An alternative statistical measure of Langmuir circulation Finally, it can be remarked that Langmuir turbulence orientation is the direction determined by a peak in the distribution 2 differs from shear turbulence in that the negative feed- function for tan 1(vy/vx), where v is the vorticity vector (see, e.g., back between the turbulent kinetic energy production and Van Roekel et al. 2012). This gives similar results to the correlation P › h i function method used here, but with somewhat greater estimation the mean shear is not as tight: u is proportional to z u , noise. which in the absence of waves will be reduced if e and 8 The elliptical fit parameters are only weakly dependent on the P «LC momentum mixing increase, whereas st is proportional to Cw value as long as is not too close to 1. MARCH 2014 M C W I L L I A M S E T A L . 885 swell aligned with the wind-sea, uLC(z) is nearly wind and south cases. Furthermore, the Langmuir circula- aligned at the surface and progressively rotates clock- tion pattern complexity in Fig. 11 is not always well wise with depth, while «LC(z) is largest at the surface and characterized by only a single angle, even at a fixed decreases with depth down to near isotropy at the bot- depth. Therefore, uLC does not seem to be a robust tom of the boundary layer. Neither of these behaviors is property of Langmuir turbulence with different swell- very different due to the presence of swell. However, wave orientations. when the swell is misaligned with the wind-sea, the The Langmuir circulation pattern dependencies are Langmuir circulations rotate in the direction of the swell not strong with Ds variations in case set Ds (again not by a small amount near the surface. Below the surface, shown), except for a moderate enhancement of the w the behavior of uLC(z) is very different for us 56p/2: it magnitudes near the surface in the case with the smallest rotates clockwise even more rapidly for southward swell, Ds value of 4 m (i.e., a steep swell wave due to its short and it is nearly independent of depth for northward wavelength of l 5 50 m). swell. Additionally, for northward swell, the circulation A further structural oddity emerges in the limit of anisotropy is very much enhanced in the middle of the large swell and weak wind (i.e., in the case set Ua for the 21 layer, compared to the surface and compared to the smallest Ua value of 2.5 m s ; Fig. 12). The Langmuir other us orientations. circulations become much larger in scale, both in their Beyond these statistical measures, the Langmuir cir- along-axis extent and in the spacing between neighbor- culation patterns are displayed in Fig. 11 for visual as- ing circulations. In this case, they are bundles of aniso- sessment. Near the surface (z 521 m) the patterns are tropic overturning cells rather than individual isolated only moderately different for different values of Us and cells. Within the bundles are multiple spatial scales us, although some variations in w amplitude and an- and orientation angles. A primary orientation—roughly isotropy direction are discernible. By the middle of the northwest (NW)–SE at both depths shown in the figure— layer (z 5210 m), however, the pattern differences are characterizes the long axis of the bundle, but different quite large. Besides simple amplitude dependencies for orientations arise on a finer scale within the bundles— w—larger w both for larger Us and for us in the range more E–W near the surface and more N–S in the boundary from SE to N—the nonmonotonic dependency of uLC on layer interior. This is certainly a highly organized turbu- us and the highly variable degrees of pattern complexity lent coherent structure, but it is a long way from an ide- are notable. For Us 5 0, the pattern is a familiar one alized roll cell. For the smallest La values in Fig. 9, some from previous studies of equilibrium Langmuir turbu- sort of regime transition is occurring, with a break from 2 lence. For an opposing swell (us 5 p), the Langmuir the ;La 4/3 scaling growth of hw2i in the boundary layer, circulations are regular roll cells with nearly uniform while the scaling growth of hw2i below the boundary layer spacing and a north–south orientation perpendicular to continues, as does the growth of hwTi (Fig. 9). Any both the wind-sea and the swell-sea. But with the swell in changes in the growth of e are less clear. the SE–N sector, there are multiple Langmuir circula- Comparison of the statistics and patterns among the tion orientations and scales and complicated branching cases in set Lh indicate that there is only a weak de- patterns from a dominant w extremum. It is clear that no pendency on the domain width L, with the widest domain simple conception of the structure of Langmuir circu- (L 5 1200 m) large enough to encompass several of the lations holds across the range of swell conditions. large-scale Langmuir circulations, suggesting a conver- In Van Roekel et al. (2012), considerable attention is gence of the LES solutions with this essentially compu- given to the near-surface orientation angle of Langmuir tational parameter. If solutions were sought with even circulations uLC as the misalignment angle between an smaller La values (e.g., even weaker winds), it is likely equilibrium wind-sea and the wind stress is varied over that the Langmuir circulation size would grow even a range from um 5 0to13p/4. The principal motivation larger, requiring larger L. Malecha et al. (2013) have is to interpret the dependency of hw2i on the wave pa- proposed a multiscale asymptotic treatment in the limit of rameters (further discussed in section 8). Their result is La / 0 as an aid to computational efficiency; it is an open that uLC monotonically increases with um (their Figs. 2 question whether this is a viable approach to dealing with and 7). In that paper several statistical predictors of uLC the complex bundle structures evident in Fig. 12. are assessed, with varying degrees of skill, and mostly uLC is about half way between the wind and wave angles 8. Summary and discussion for this set of cases. In Fig. 10, the sign of uLC near the surface is the same as the sign of us, but its magnitude is The presence of remotely generated swell waves sig- much less than us/2. In the boundary layer interior, the nificantly alters the Langmuir turbulence of an equilib- uLC(z) profile shapes are quite different in the north rium wind-sea. The Lagrangian-mean flow profile uL(z) 886 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 44

FIG. 11. The w(x, y) snapshots at z 5 (top) 21 and (bottom) 210 m, normalized by u . In the center is the case from the set Us1 with no swell (Us 5 0), and around the * 2 edge are the cases in set Qs(Us 5 0.25 m s 1, Ds 5 16 m) located in their appropriate compass positions. MARCH 2014 M C W I L L I A M S E T A L . 887

not so complete as to leave uL unaltered with added swell-waves. The Reynolds stress profile R(z) also has a fatter Ekman spiral with strong swell, and with mis- alignment it rotates to a moderate degree toward the swell direction within the boundary layer interior. For a given swell magnitude, the orientation of opposing wind- and swell- (i.e., us 56p) yields the weakest uL and thinnest Ekman spiral. All of these behaviors are similar to the analytic Ekman–Stokes model solutions with constant density and eddy viscosity (section 3). Turbulent intensity, e and hw2i, and entrainment rate hwTi are inverse functions of Langmuir number La. Within the boundary layer the intensity measures 2 scale roughly as La 4/3, consistent with an argument based on Stokes production as the dominant source of Langmuir turbulence, while for the inertial current, entrainment rate, and fluctuation variances below the boundary layer, 2 the dependencies are even steeper, perhaps ;La 2 (as predicted for Uin 0 by the Ekman–Stokes model). The Ekman–Stokes model predicts a reflection sym- metry in the swell direction (i.e., independence of the sign of us) for the mean boundary layer, but this does not hold in the LES solutions. While uL(z) is approximately symmetric in its weakening and rotation for positive and negative us (Fig. 3, right), most other boundary layer properties are asymmetric. This is strongly so for inertial current amplitude, turbulent intensity, and entrainment rate, all of which are much stronger with leftward (us . 0; counterclockwise, anticyclonic, and northward) swell misalignment than with rightward misalignment. The Langmuir circulation patterns are similarly asymmetric with us, with stronger, more vertically aligned, and more elongated eddies when us . 0.9 As La decreases with larger swell, the Langmuir circulation patterns become more complex and multiscale; at the smallest value of La ’ 0.1 included here, the outermost longitudinal and transverse correlation scales become very large com- pared to the boundary layer depth (Fig. 12). Even with a large added swell, however, the general characteristics of Langmuir turbulence are similar to FIG. 12. The w(x, y) snapshots at z 5 (top) 21 and (bottom) what occurs in the equilibrium wind-sea regime, but the 2 5 21 s 5 21 10 m for the case set Lh (i.e., Ua 2.5 m s , U 0.25 m s , quantitative differences are not small. In the swell-wave s s u 5 0, D 5 16 m, and variable L) with La 5 0.11. The full panel is parameter space of (Us, us, Ds), the most important pa- for the quadrupled horizontal domain case with L 5 1200 m, and rameter is La in (3). With aligned swell, as Us increases the lower-left inset is for the standard domain with L 5 300 m at an uncorrelated time. and La decreases, the Lagrangian-mean flow, turbulent intensity both within the boundary layer and below, and has a larger surface value and a fatter Ekman spiral entrainment rate all increase. As us varies away from throughout the boundary layer with increasing swell wind alignment, La in (3) increases due to swell- and magnitude Us. With misaligned wind and swell, uL ro- tates toward the swell direction, but only within a limited 6 8 angular range that is almost equal to 30 . The Eulerian- 9 The nonequilibrium misalignment cases examined in Van mean flow hui(z) exhibits an anti-Stokes component that Roekel et al. (2012) all have nonnegative wave rotation angles becomes large when Us is large, but this cancellation is relative to the wind u equal to or greater than zero. 888 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 44 wind-sea Stokes drift misalignment. This increase in La complexity of the swell influences on the eddy viscosity accounts partly for the general tendencies for decreases profile diagnosed from the LES solutions [i.e., kL(z)in s in the preceding quantities; however, the u dependency Fig. 6] indicates that neither the constant ko in the is asymmetric in its sign in the LES (unlike in the Ekman–Stokes model nor the [K(z), 0] in the KPP scheme Ekman–Stokes model or in the La definition), and would yield a highly accurate uL(z). In particular, the sometimes the asymmetry is strong enough to overcome magnitude of kL increases in equilibrium wind-wave the implied change due to La (e.g., the peak value for Langmuir turbulence compared to shear turbulence with- e averaged over the boundary layer occurs at us 5 p/2 in out waves (McWilliams and Sullivan 2000; McWilliams Fig. 8). The Ds dependencies are mostly rather weak, et al. 2012), but as La further decreases with swell waves, although inertial amplitude Uin 0 and diagnosed eddy the eddy viscosity magnitude decreases (section 4); that is, viscosity kL(z) are exceptions. This is consistent with the jkj must have a maximum at an intermediate value of La absence of any Ds influence in the definition of La in (3), near its wind-sea equilibrium value of 0.3. These results which differs from the alternative definition in Harcourt raise a subtle design issue about the trade-off between and D’Asaro (2008) in this aspect. parameterization simplicity (which often correlates with In Van Roekel et al. (2012, Fig. 12) functional fits robustness across regimes) and hui(z) profile accuracy are presented for hw2i from LES cases as alternative from the perspective of practical utility in circulation functions of alternatively defined Langmuir numbers models. This decrease in jkj for very small La happens even that combine the influences of Stokes drift magnitude, while e and hw2i continue to increase with decreasing La direction, and depth scale. The latter alternatives in- because of increasing turbulent kinetic energy production clude different depth weighting for the Stokes drift (section 6); these opposing tendencies contradict the mix- magnitude in the denominator [as originally proposed ing length representation of k ; wsh,withws as a turbulent by Harcourt and D’Asaro (2008) to exclude de- velocity scale. Our present view is that relatively simple pendencies on the near-surface wind-sea ust magnitude rules for the dependency of k(z) on the wave parameters and on its extension beyond the boundary layer when Ds . (i.e., simpler than manifested in Fig. 6) may provide a useful h] and different angular projections for the misaligned accuracy for hui(z). swell, wind, and Langmuir circulations. While all of the For hTi(z) and other material properties, the impor- fits are somewhat skillful, the skill levels among the al- tant parameterization target of Langmuir turbulence is ternatives are not sharply distinguished. The swell in- the entrainment rate (Fig. 9). Its dependencies on the fluences summarized in the preceding paragraph do not swell parameters are fairly simple, and they mirror those accurately collapse into a universal function of any of of the depth-averaged turbulent intensity (Fig. 8). The these alternative Langmuir numbers (not shown). In entrainment rate, represented in the KPP scheme as the LC particular, the orientation angle u is a complex function negative peak in hwTi(z) 52Ks(z)›zhTi(z) near the top of us and depth (Figs. 10 and 11); hence, it is not useful as of the , is a sensitive function of the eddy an a priori projection angle for a composite La. diffusivity profile near the bottom of the boundary layer, Oceanic circulation models require a boundary layer most importantly how far the diagnosed boundary layer parameterization because their computational grids are depth at z 52h penetrates into the stable stratification. necessarily too coarse to resolve Langmuir turbulence. The determination of h is represented in the KPP With data inputs about the surface fluxes and gravity scheme for the free convection regime by a critical bulk wave field, the desired outcome is accurate profiles of Richardson number condition where the mean velocity L 10 hui(z) [or u (z)], hTi(z), and other scalar profiles de- scale is augmented by a turbulent velocity scale Vt. The termined in competition with larger-scale dynamical influences, while accurate characterizations of turbulent 10 quantities such as e, hw2i, and k(z) are only means to that In present implementations of the KPP scheme in circulation end. In particular, the swell influences demonstrated models, there are two alternative formulations for the critical Richardson number condition that determines the boundary layer here do need to be parameterized. depth h. One formulation is expressed as a ratio of the bulk dif- L For u (z) and the inertial current amplitude, even the ferences across the boundary layer of buoyancy change and ve- 2 Dh 2i simple Ekman–Stokes model with constant eddy vis- locity change squared; in this representation Vt is added to u in cosity is fairly skillful (Figs. 3, 4, and 5), even though in the denominator (Large et al. 1994, section 3). The other formu- practice the parameterization must also account for the lation is expressed as a boundary layer integral condition of the › h i 2 influence of stratification in limiting the boundary layer competition between the destabilizing influence of ( z u ) and the stabilizing influence of positive buoyancy stratification ag›zhTi;in depth and the generally convex shape of the eddy vis- 2 this representation, Vt /h is added to the vertical integral of the cosity profile [both of which are parts of the KPP mean shear variance (Lemarie et al. 2012, section 3). Both for- scheme; Large et al. (1994)]. On the other hand, the mulations have essentially the same dynamical rationale. MARCH 2014 M C W I L L I A M S E T A L . 889 rationale is that entrainment by encroachment into the In conclusion, the two relevant idealized regimes of thermocline can happen even when the mean shear equilibrium Langmuir turbulence in the —with is weak, as in convection at low wind speed or as in both wind and waves held steady in time and the tur- Langmuir turbulence with strong waves and weak winds. bulence evolving into a stationary state—are for a wind-

To choose Vt, one must specify the desired entrainment sea alone, as previously analyzed in many papers, and rate that will result from the KPP scheme when Vt is for the addition of a remotely generated swell-sea, as large enough to dominate the mean shear in determining analyzed here. Beyond these, the relevant regimes are h. For free convection, the entrainment rule is the peak inherently transient in both the winds and waves, and value of the pycnocline buoyancy flux, which is a nega- often the turbulence as well. Only a few transient situ- tive fraction, bT ’20.15, of the destabilizing surface ations have been addressed, for example, spinup from upward buoyancy flux B . 0. For Langmuir turbulence rest, the passage of a hurricane utilizing a transient, with swell, the entrainment rule is expressed by the nonlocal wave model (Sullivan et al. 2012), and com- normalized functional dependencies plotted as the ordi- bined measurements of variable waves and currents nates in Fig. 9, which are denoted as F(Us, us). With these (Smith 1992, 1998). Therefore, to make further progress rules a combined expression can be derived for Vt with it is of pressing importance to measure and model across both convection and Langmuir turbulence, following the a wide range of transient situations to discover how they arguments in Large et al. (1994, p. 372); namely, differ from equilibrium Langmuir turbulence. " # 1/2 Acknowledgments. This research was sponsored by the C Nhw (2b )w3 1 2u3 F V2 5 y s T * * . (28) National Science Foundation (Grant DMS-785 0723757) t Ri w3 cr s and Office of Naval Research (Grant N00014-08-1-0597). Computations were performed on the supercomputers The notation is as in that paper: N is the buoyancy fre- Bluefire at NCAR and Thresher, Trestles, and Gordon at quency just below the boundary layer; ws is the turbulent SDSC. We appreciate Dr. Leonel Romero’s guidance on velocity scale used in scaling the eddy diffusivity for the theory and observations of swell-waves and his help in ; material , Ks(z) wsh;Ricr is the critical preparing Fig. 1. value for the bulk Richardson number; w ; (Bh)1/3 is * the conventional convective velocity scale determined REFERENCES from the surface buoyancy flux and boundary layer depth; and Cy is an empirical coefficient slightly larger Alves, J. H. G. M., 2006: Numerical modeling of ocean swell con- than the one related to the shape of hTi(z) in the en- tributions to the global wind-wave climate. Ocean Modell., 11, trainment layer. The first term in the second numerator 998–112. ——, M. L. Banner, and I. R. Young, 2003: Revisiting the Pierson– is the previously proposed convective contribution, and Moskowitz asymptotic limits for fully developed wind waves. the second term is the new Langmuir contribution. In- J. Phys. Oceanogr., 33, 1301–1323. terpreting the result of Fig. 9 (left) as an experimental Belcher, S. E., 2012: A global perspective on Langmuir turbulence scaling relation for the normalized entrainment rate of in the ocean surface boundary layer. Geophys. Res. Lett., 39, 2 F ; La 2, the scaling dependency of V from (28) in L18605, doi:10.1029/2012GL052932. t Craik, A. D. D., and S. Leibovich, 1976: Rational model for Langmuir turbulence is Langmuir circulations. J. Fluid Mech., 73, 401–426. Depley, M. T., F. Ardhuin, F. Collard, and B. Chapron, 2010: ; 1/2 21/3 Space-time structure of long ocean swell fields. J. Geophys. Vt (Nhu ) La , (29) * Res., 115, C12037, doi:10.1029/2009JC005885. Grant, A. L. M., and S. E. Belcher, 2009: Characteristics of Lang- ; 22/3 assuming that ws u* La (i.e., the scaling curve in muir turbulence in the ocean mixed layer. J. Phys. Oceanogr., Fig. 8). 39, 1871–1887. The parameterization discussion in the preceding two Hanley, K. E., S. E. Belcher, and P. P. Sullivan, 2010: A global paragraphs sketches how the KPP scheme could be climatology of wind-wave interaction. J. Phys. Oceanogr., 40, 1263–1282. modified to include the important effects of both a wind- Harcourt, R. R., 2013: A second-moment closure model of Lang- sea and swell in Langmuir turbulence by generalizing Vt muir turbulence. J. Phys. Oceanogr., 43, 673–697. in the condition determining h and by choosing the ——, and E. A. D’Asaro, 2008: Simulation of Langmuir turbulence J. Phys. Oceanogr., turbulent velocity scale ws to have an appropriate de- in pure wind seas. 38, 1542–1562. pendency on the wave parameters La and us (and per- Large, W. G., J. C. McWilliams, and S. Doney, 1994: Oceanic s vertical mixing: A review and a model with a nonlocal boundary haps D ). However, a complete and implementable layer parameterization. Rev. Geophys., 32, 363–403. parameterization scheme requires extensive testing, and Leibovich, S., 1983: The form and dynamics of Langmuir circula- that task is beyond the scope of this paper. tions. Annu. Rev. Fluid Mech., 15, 391–427. 890 JOURNAL OF PHYSICAL OCEANOGRAPHY VOLUME 44

Lemarie, F., J. Kurian, A. F. Shchepetkin, M. J. Molemaker, Smith, J. A., 1992: Observed growth of Langmuir circulation. F. Colas, and J. C. McWilliams, 2013: Are there inescapable J. Geophys. Res., 97 (C4), 5651–5664. issues prohibiting the use of terrain-following coordinates in ——, 1998: Evolution of Langmuir circulation during a storm. climate models? Ocean Modell., 42, 57–79. J. Geophys. Res., 103 (C6), 12 668–12 694. Malecha, Z., G. Chini, and K. Julien, 2013: A multiscale algorithm Sullivan, P. P., and J. C. McWilliams, 2010: Dynamics of winds and for simulating spatially-extended Langmuir circulation dy- currents coupled to surface waves. Annu. Rev. Fluid Mech., 42, namics. J. Comput. Phys., in press. 19–42. Masson, D., 1993: On the nonlinear coupling between swell and ——, ——, and W. K. Melville, 2007: Surface gravity wave effects in wind waves. J. Phys. Oceanogr., 23, 1249–1258. the oceanic boundary layer: Large eddy simulation with vortex McWilliams, J. C., and P. P. Sullivan, 2000: Vertical mixing by force and stochastic breakers. J. Fluid Mech., 593, 405–452. Langmuir circulations. Spill Sci. Technol. Bull., 6, 225–237. ——, J. B. Edson, T. Hristov, and J. C. McWilliams, 2008: Large- ——, ——, and C. H. Moeng, 1997: Langmuir turbulence in the eddy simulations and observations of atmospheric marine ocean. J. Fluid Mech., 334, 1–30. boundary layers above nonequilibrium surface waves. J. At- ——, E. Huckle, J. Liang, and P. Sullivan, 2012: The wavy Ekman mos. Sci., 65, 1225–1245. layer: Langmuir circulations, breakers, and Reynolds stress. ——, L. Romero, J. C. McWilliams, and W. K. Melville, 2012: J. Phys. Oceanogr., 42, 1793–1816. Transient evolution of Langmuir turbulence in ocean bound- Polton, J. A., J. A. Smith, J. A. MacKinnon, and A. E. Tejada- ary layers driven by hurricane winds and waves. J. Phys. Martinez, 2008: Rapid generation of high-frequency internal Oceanogr., 42, 1959–1980. waves beneath a wind and wave forced oceanic surface Van Roekel, L. P., B. Fox-Kemper, P. P. Sullivan, P. E. Hamlington, mixed layer. Geophys. Res. Lett., 35, L13602, doi:10.1029/ and S. R. Haney, 2012: The form and orientation of Langmuir 2008GL033856. cells for misaligned wind and waves. J. Geophys. Res., 117, Price, J. F., 1981: Upper ocean response to a hurricane. J. Phys. C05001, doi:10.1029/2011JC007516. Oceanogr., 11, 153–175. Young, I. R., 1999: Wind Generated Ocean Waves. Elsevier, 287 pp. Romero, L., and W. K. Melville, 2010: Airborne observations of Zakharov, V. E., 2005: Theoretical interpretation of fetch-limited fetch-limited waves in the Gulf of Tehuantepec. J. Phys. wind-driven sea observations. Nonlinear Processes Geophys., Oceanogr., 40, 441–456. 12, 1011–1020.