1. Introduction
Studies of the upper ocean advance the understanding necessary to develop rules relating the turbulence to the forcing and the ocean state. These mixing rules allow turbulent fluxes to be estimated from information provided by observations, numerical models, or a combination of both through data assimilation. Unfortunately, the technology to measure fluxes directly in the open ocean does not yet exist, so observational knowledge is limited to indirect inferences, such as those from ocean microstructure (e.g., Moum et al. 2013).
In ocean general circulation models vertical mixing schemes employ a set of established rules to parameterize the vertical fluxes, whose divergences appear as terms in the prognostic equations for the resolved ocean state. Such schemes come in two distinct flavors, that Burchard et al. (2008) term statistical turbulence models (STM) and empirical turbulence models (ETM). STMs solve the Reynolds-averaged Navier–Stokes equations with a variety of second-order closures (e.g., Mellor and Yamada 1982; Kantha and Clayson 1994; Harcourt 2015) and some higher order (e.g., Cheng et al. 2002; Canuto et al. 2009). Empirical relationships simplify ETMs, such as the mixed-layer models of Kraus and Turner (1967) lineage, and of Price et al. (1986) and the K-profile parameterization (KPP) of the ocean boundary layer (Large et al. 1994). In the absence of direct measurements, evaluation of all schemes, as well as guidance on further developments, have relied heavily on comparisons of modeled and observed ocean states. However, discrepancies are caused both by inadequate rules and by forcing error, with no means to discriminate and the possibility of compensating errors a further complication.
An extensive comparison by Li et al. (2019) of 11 vertical mixing schemes, including both STM and ETM, demonstrates the current lack of community consensus. The primary metric is the mixed layer depth (MLD) of de Boyer Montégut (2004), and a robust conclusion is that the inclusion of Langmuir turbulence produces deeper MLD. Although this measure would be insufficient for either showing comprehensive agreement or for judging relative merit, it does expose “limited understanding and numerical deficiencies.” In particular, no two schemes agree globally over the annual cycle, so overall they do not perform as well as might be expected based on the theory, observations, and numerical experiments applied to their development. One of the focused demonstrations of disparate behavior is the annual cycle at Ocean Weather Station Papa, but unbalanced surface fluxes preclude the quantitative evaluations against observations of both the forcing and ocean state found in Martin (1985) and Large (1996), for example.
Large-eddy simulation (LES) is fundamentally different and offers unique opportunities, because it can be used to study turbulence (Burchard et al. 2008). LES models solve the Navier–Stokes equations by resolving the vertical fluxes down to scales of order a meter, such that subgrid-scale (SGS) contributions are small and relatively well parameterized. LES is key to many recent advances in the understanding and hence modeling of upper-ocean mixing physics (e.g., McWilliams et al. 1997; Wang et al. 1998; Grant and Belcher 2009; Harcourt and D’Asaro 2008; van Roekel et al. 2012; Li and Fox-Kemper 2017). The present study of the diurnal cycle follows directly from two particular examples: Large et al. (2019a, hereafter L19a) extends Monin–Obukhov similarity theory (Monin and Obukhov 1954) to include surface wave forcing (Stokes drift), and Large et al. (2019b, hereafter L19b) relates this forcing to both the local and the nonlocal transports of buoyancy and momentum throughout the boundary layer. Hereafter, these works will collectively be referred to as L19. These results and additional analyses of the same LES data (Large et al. 2021) are herein combined to develop rules for boundary layer entrainment and detrainment and for misalignment of the shear and turbulent stress vectors when changes in the forcing may be too rapid for equilibrium assumptions.
The LES model, the surface flux and wave forcing, and the various simulations are detailed L19, so only the most relevant aspects are recapped in section 2. Germane results from L19 are summarized in section 3, where there is also a novel analysis of similarity theory in the stably forced surface layer. In section 4, an empirical rule that includes Stokes forcing is developed for entrainment into both convective nighttime and stable afternoon boundary layers. In addition, the orientation of the stress vector is related to the Eulerian and Lagrangian shear over different vertical scales. Section 5 focuses on detrainment, which is broadly defined as daytime turbulence that depends strongly on past history, such as found below a stable morning boundary layer. In section 6, the diurnal cycles of buoyancy and momentum fluxes in the upper ocean are parameterized in two distinct ways, then both are evaluated against LES fluxes to take advantage of the LES mean ocean state and turbulence statistics being consistent with the specified “error-free” forcing.
2. LES of the Southern Ocean boundary layer
Large-eddy simulations of the ocean boundary layer are driven by the observed forcing shown in Fig. 1 at the Southern Ocean Flux Site (SOFS) near 47°S, 140°E (Schultz et al. 2012). As detailed in L19a, there are two cases from April 2010: AprS with wave forcing from realistic Stokes drift profiles and AprN without. Two additional cases from June 2010 are similarly denoted as JunS and JunN. Also utilized here is an idealized case, D24S, where the surface waves are always perfectly aligned with a westerly wind. All cases use the SOFS inertial period of 16.4 h. This study of the diurnal cycle focuses on AprS, because it is the only Stokes case with a period of surface buoyancy gain (Fig. 1c), beginning with the transition at tUS = 32.75 h from unstable buoyancy forcing (effective cooling) to stable (heating), until the reverse at tSU = 40.75 h. The other cases extend the range of forcing conditions, including the prior April diurnal cycle, as well as another two in June (noon at hours 6 and 30) when there is net cooling throughout.

SOFS surface forcing for April (red) and June (blue), and the idealized forcing of D24S (gray): (a) friction velocity
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

SOFS surface forcing for April (red) and June (blue), and the idealized forcing of D24S (gray): (a) friction velocity
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
SOFS surface forcing for April (red) and June (blue), and the idealized forcing of D24S (gray): (a) friction velocity
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
The LES model is well documented (Moeng 1984; Sullivan et al. 1994; McWilliams et al. 1997) and widely used in ocean boundary layer studies (e.g., Sullivan et al. 2012; van Roekel et al. 2012; Kukulka et al. 2013). The modifications to account for surface wave effects are extensive and summarized in L19a, as are the initial conditions and grid choices. Initially, the inversion depth of maximum stratification di is about 180 m in April, 222 m in June, and 96 m for D24S. The model solves the wave-averaged, incompressible, and Boussinesq Craik–Leibovich equation set (Craik and Leibovich 1976; McWilliams et al. 1997). Its thermodynamic variable is buoyancy g(1 − ρ/ρo), relative to a reference density ρo, where g is gravitational acceleration and ρ is ocean density. It produces evolving profiles of horizontal and time mean buoyancy Θ and horizontal Eulerian flow U, with orthogonal components, U and V. The corresponding turbulent vertical fluxes, ⟨wθ⟩, ⟨wu⟩, and ⟨wυ⟩, are given by the correlations of vertical velocity w with fluctuations, θ, u, and υ, plus small SGS contributions. Similarly, the variance of velocity fluctuations plus SGS give the turbulent kinetic energy, TKE, whose dissipation εTKE is parameterized. Hourly statistics converge and are computed half-hourly in order to track turbulent responses to the variable forcing that are faster than the responses of stratification and shear, and hence of eddy diffusivity and viscosity. Thus, only every second calculation is independent.
a. Meteorological and Stokes (wave) forcing
Surface waves are imposed as profiles of Stokes drift
Figure 2 shows the AprS distributions below 1 m of TKE, its dissipation, and its three production terms of both signs. The column integrals of these three terms above the inversion depth (Fig. 2d) are displayed as time series in Fig. 1. The buoyancy integral is often small (Fig. 1f), because of significant TKE suppression (⟨wθ⟩ < 0), and can become negative with heating (Q0 > 0). Usually the column Eulerian production is about

The AprS nondimensional distributions below 1 m of (a) TKE and its (b) buoyancy production, (c) Eulerian shear production, (d) Stokes shear production, and (e) dissipation. The contours are irregular at ±0.10, ±0.30, ±1.0, ±3.0, ±10.0, and 30, with negative contours gray. The time series (red) show the boundary layer depth hL19 (solid), the entrainment depth dE (dotted line), the inversion depth di (long-dashed line) of maximum stratification, and
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

The AprS nondimensional distributions below 1 m of (a) TKE and its (b) buoyancy production, (c) Eulerian shear production, (d) Stokes shear production, and (e) dissipation. The contours are irregular at ±0.10, ±0.30, ±1.0, ±3.0, ±10.0, and 30, with negative contours gray. The time series (red) show the boundary layer depth hL19 (solid), the entrainment depth dE (dotted line), the inversion depth di (long-dashed line) of maximum stratification, and
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
The AprS nondimensional distributions below 1 m of (a) TKE and its (b) buoyancy production, (c) Eulerian shear production, (d) Stokes shear production, and (e) dissipation. The contours are irregular at ±0.10, ±0.30, ±1.0, ±3.0, ±10.0, and 30, with negative contours gray. The time series (red) show the boundary layer depth hL19 (solid), the entrainment depth dE (dotted line), the inversion depth di (long-dashed line) of maximum stratification, and
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
b. Diagnostic depths, regimes, and layers
Surface forced turbulence, including Stokes effects, dominates at depths, d = −z, in the boundary layer where σ = d/h is a natural vertical coordinate. Table 1 catalogs various boundary layer depths as well as other diagnostic and empirical depths and angles. In L19, the buoyancy flux (e.g., Fig. 2b) is used to diagnose the entrainment depth (dE, dotted red trace) where the entrainment buoyancy flux BE is a negative minimum, as well as a deeper boundary layer depth hL19 (solid red traces) where the flux approaches a local maximum, which is usually near zero. However, when the surface forcing does not dominate, this maximum can be shallow and significantly positive, such that the boundary layer and interpretation of hL19 become ambiguous. In particular, Fig. 2b shows a very abrupt collapse of hL19 at tUS, followed by two hours of deepening then 2 h of shoaling, but this behavior is reflected in neither TKE (Fig. 2a), nor its dissipation (Fig. 2e).
Catalog of depths and of angles between horizontal vectors, including empirical relationships, brief descriptions, basic dependencies, and references.


The examples of Fig. 3 show profiles of buoyancy flux (blue traces) and smoothed stratification (red traces). The stratification is scaled with depth, so it is constant where the buoyancy profile is logarithmic. The additional scaling by

Profiles of buoyancy flux (blue) normalized by the greater of |B0| or |BE|, and of smoothed stratification (red) scaled (top) by
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

Profiles of buoyancy flux (blue) normalized by the greater of |B0| or |BE|, and of smoothed stratification (red) scaled (top) by
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Profiles of buoyancy flux (blue) normalized by the greater of |B0| or |BE|, and of smoothed stratification (red) scaled (top) by
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Both entrainment regimes are characterized by buoyancy loss (∂z⟨wθ⟩ > 0) between the surface and dE and by gain below as denser water is mixed into the boundary layer. Regime EC covers convective entrainment into strong stratification near the inversion. In the typical example of Fig. 3a, stratification is unstable above 80 m and the nearly logarithmic buoyancy profile is consistent with similarity theory even beyond the surface layer. Also, ⟨wθ⟩ is countergradient between 75 and 100 m. Apart from the WRW period, Fig. 2 shows that TKE scales well with
The onset of solar radiation, with weak winds initiates a transition from regime EC to DBL at time, tED = 30.5 h of AprS. Regime DBL has unstable forcing (B0 < 0) to tUS and lasts until a transition to regime SBL at about tDS = 35.5 h. The profiles of Fig. 3b are representative of the entire regime, including a period of JunS (Fig. 1; dotted blue). The robust characteristics of detrainment are negligible entrainment; negative curvature of the buoyancy flux profile in a detrainment zone from the surface to near the entrainment depth at the transition [dE(tED) about 155 m]; hL19 approaches the depth dMAX where the flux is a positive maximum denoted as BMAX; no logarithmic layer above hL19; and extensive regions of countergradient buoyancy flux both above and below dMAX. In Fig. 2b, dMAX steadily deepens from the surface at tED to about 100 m at tDS, while
The Fig. 3c example from regime SBL (from tDS until the transition to regime EW at tSE = 38.5 h) appears to be a near equilibrium stable boundary layer, because there is no evidence of either entrainment (dE = 0), or detrainment (BMAX ≤ 0). Its characteristic depth is relatively shallow, with
3. The upper-ocean boundary layer
The nondimensional shape functions Gm,s are found to be quite similar in L19b, and the combined averages over 20 bins are shown in Fig. 4 (

Nondimensional shape functions of (top) σ > 0.5 and (bottom) σ < 0.5: the revised composite GC (solid black traces); and the composite bin averages
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

Nondimensional shape functions of (top) σ > 0.5 and (bottom) σ < 0.5: the revised composite GC (solid black traces); and the composite bin averages
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Nondimensional shape functions of (top) σ > 0.5 and (bottom) σ < 0.5: the revised composite GC (solid black traces); and the composite bin averages
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
The surface layer

Similarity functions of the stability parameter (14) calculated from hourly statistics every half hour during the stable buoyancy forcing of AprS for (a) momentum, with ϕm = 1 + 14ζ (dashed) and (b) buoyancy, with ϕs = 1 + 5ζ (dashed).
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

Similarity functions of the stability parameter (14) calculated from hourly statistics every half hour during the stable buoyancy forcing of AprS for (a) momentum, with ϕm = 1 + 14ζ (dashed) and (b) buoyancy, with ϕs = 1 + 5ζ (dashed).
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Similarity functions of the stability parameter (14) calculated from hourly statistics every half hour during the stable buoyancy forcing of AprS for (a) momentum, with ϕm = 1 + 14ζ (dashed) and (b) buoyancy, with ϕs = 1 + 5ζ (dashed).
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
4. Entrainment
A fraction of the turbulent kinetic energy produced in a boundary layer is not dissipated, but converted into column potential energy when denser water is entrained at depth then mixed throughout. An empirical rule for atmospheric convection says that the ratio of entrainment to surface buoyancy flux (BE/B0) is constant (Ball 1960) at about 0.2 (Tennekes 1973). An interpretation of this rule is that the integrated buoyancy source (⟨wθ⟩ > 0) of TKE is proportional to
Since entrainment (BE < 0) requires stable stratification near the entrainment depth, there should be a dependence on a function F(δρ) of a bulk density difference, δρ > 0. For present purposes F(δρ) is unity, until a much greater range of ocean conditions shows how it approaches zero for an unstratified water column, and possibly increases for entrainment into the much stronger and shallower stratification of a seasonal pycnocline, for example.

The entrainment rule from independent hourly statistics showing the product −BEdE vs (a) net surface layer production of TKE
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

The entrainment rule from independent hourly statistics showing the product −BEdE vs (a) net surface layer production of TKE
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
The entrainment rule from independent hourly statistics showing the product −BEdE vs (a) net surface layer production of TKE
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
A similar procedure can relate any measure of entrainment to the source integrals over different depth ranges. For example, ME, the integral of (4) over the entrainment zone of ⟨wθ⟩ < 0, gives the rate of conversion of TKE to potential energy across the zone. It is the basis for determining the all-important boundary layer depth in the ePBL scheme of Reichl and Hallberg (2018) and its extension to Langmuir turbulence by Reichl and Li (2019). Its correlation coefficient with
Another application of the procedure relates −BEdE to the source integrals above −hL19, −dE, or −di. With any of these limits, the resulting regression coefficients corresponding to those of (17) become 0.020, 0.035, and 0.27. Respectively, these values are the fractions of TKE produced by Eulerian shear, Stokes shear, and buoyancy that go to increasing potential energy rather than dissipating. Buoyancy is by far the most efficient with over 25% going to entrainment, and so contributes about 10% of the regime EC driving of AprS before the winds weaken at hour 15, and as much as 90% afterward, but is negligible in regime EW. Although much less efficient, the Stokes source usually contributes more than 70% of the entrainment driving, except for the WRW period. The largest Eulerian shear contribution of AprS is only about 20% in regime EW.
a. Entraining boundary layer depth

Comparison of the boundary layer depth hBL to other Table 1 depths diagnosed from 224 independent hourly statistics throughout the entrainment regimes of AprS, JunS, and D24S with Stokes forcing (red) and from AprN and JunN (blue): (a) entrainment depth dE, (b) hL19, (c) the KPP boundary layer depth of Danabasoglu et al. (2006), and (d) parameterized hER from the entrainment rule (section 6).
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

Comparison of the boundary layer depth hBL to other Table 1 depths diagnosed from 224 independent hourly statistics throughout the entrainment regimes of AprS, JunS, and D24S with Stokes forcing (red) and from AprN and JunN (blue): (a) entrainment depth dE, (b) hL19, (c) the KPP boundary layer depth of Danabasoglu et al. (2006), and (d) parameterized hER from the entrainment rule (section 6).
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Comparison of the boundary layer depth hBL to other Table 1 depths diagnosed from 224 independent hourly statistics throughout the entrainment regimes of AprS, JunS, and D24S with Stokes forcing (red) and from AprN and JunN (blue): (a) entrainment depth dE, (b) hL19, (c) the KPP boundary layer depth of Danabasoglu et al. (2006), and (d) parameterized hER from the entrainment rule (section 6).
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Figure 7c compares the KPP boundary layer depth calculation, hKPP, as updated in Danabasoglu et al. (2006). The major differences are during regime EW when hKPP becomes greater than hL19 > hBL, because it is less affected by the weak diurnal stratification near
b. Flux profiles during entrainment
The defining characteristic of the entrainment regimes is buoyancy loss (∂z⟨wθ⟩ > 0) between the surface and the entrainment depth, with gain below to beyond the boundary layer depth. Buoyancy loss through the surface (B0 < 0) is usual (Fig. 3a), but not necessary (Fig. 3d). Furthermore, the buoyancy flux varies smoothly with depth, as also seen in Fig. 8 (upper panels; solid profiles). Positive curvature across the boundary layer is a singular feature of regime EW, when the diurnal stratification weakens, and its local maximum deepens steadily from

Vertical profiles during the entrainment regimes EC (hours 9 and 26), and EW (hours 39 and 44) from AprS: (top) buoyancy flux normalized by the greater of |B0| or |BE| (solid blue), and the parameterizations from section 6, ⟨wθ⟩FP (dashed) and ⟨wθ⟩KP (dotted); (middle) stress magnitude normalized by
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

Vertical profiles during the entrainment regimes EC (hours 9 and 26), and EW (hours 39 and 44) from AprS: (top) buoyancy flux normalized by the greater of |B0| or |BE| (solid blue), and the parameterizations from section 6, ⟨wθ⟩FP (dashed) and ⟨wθ⟩KP (dotted); (middle) stress magnitude normalized by
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Vertical profiles during the entrainment regimes EC (hours 9 and 26), and EW (hours 39 and 44) from AprS: (top) buoyancy flux normalized by the greater of |B0| or |BE| (solid blue), and the parameterizations from section 6, ⟨wθ⟩FP (dashed) and ⟨wθ⟩KP (dotted); (middle) stress magnitude normalized by
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Figure 8 also shows the magnitude τ of the turbulent stress and momentum flux (middle panels; solid blue profiles) decreasing monotonically from
The Eulerian shear vector
5. The stable boundary layer and detrainment
Figure 9 shows time histories of various depths (Table 1) that bound the zones and layers of AprS (red) and AprN (blue) during the single day of stable buoyancy forcing. These depths are shown along with the profiles of ⟨wθ⟩, τ and Ωτ in the four examples of Fig. 10. The boundary layer depth of choice, h, is shown in Fig. 9a (solid traces). It follows hBL when the buoyancy forcing is unstable, and the depth, hAM (Fig. 9c), of a growing morning boundary layer from the unstable to stable transition at tUS in regime DBL until the reverse transition at tSU in regime EW.

Half-hourly time variability of diagnostic and time-integrated depths (Table 1) from hourly statistics of AprS (red) and AprN (blue) during regimes DBL, SBL, and EW: (a) hBL and the boundary layer depth of choice, h (solid traces); (b) hL19 and
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

Half-hourly time variability of diagnostic and time-integrated depths (Table 1) from hourly statistics of AprS (red) and AprN (blue) during regimes DBL, SBL, and EW: (a) hBL and the boundary layer depth of choice, h (solid traces); (b) hL19 and
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Half-hourly time variability of diagnostic and time-integrated depths (Table 1) from hourly statistics of AprS (red) and AprN (blue) during regimes DBL, SBL, and EW: (a) hBL and the boundary layer depth of choice, h (solid traces); (b) hL19 and
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

As in Fig. 8, but for the detrainment regime DBL (hours 32, 33, and 34), with
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

As in Fig. 8, but for the detrainment regime DBL (hours 32, 33, and 34), with
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
As in Fig. 8, but for the detrainment regime DBL (hours 32, 33, and 34), with
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
The boundary layer depth hL19 of Fig. 9b is always available (Fig. 2). However, a depth where the integrated Lagrangian shear production of TKE balances the buoyant suppression (gray triangles) can be found only between hours 34 and 37. In stable atmospheric boundary layers, this depth has been associated with the Monin–Obukhov depth (Wyngaard 2010), which appears to be the case near hour 35 of AprS for
a. Detrainment
Following section 2, Fig. 9d shows the depth dMAX of maximum buoyancy flux, BMAX > 0, during detrainment. Its appearance marks the transition to regime DBL at tED = 30.75 h. Such a maximum remains clearly discernible at hour 34 in Fig. 10, but not in regime SBL. In the detrainment examples of Fig. 10 (top panels), dMAX separates buoyancy gain from the loss down to near dE(tED), the entrainment depth at tED. In contrast, the corresponding stress profiles (middle panels) do not appear to reflect dMAX, but decrease monotonically from

Evaluation of the rules for regime DBL detrainment and for regime EW afternoon and nighttime entrainment from AprS (red) and AprN (blue) and from JunN and JunS (black): (a)
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

Evaluation of the rules for regime DBL detrainment and for regime EW afternoon and nighttime entrainment from AprS (red) and AprN (blue) and from JunN and JunS (black): (a)
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Evaluation of the rules for regime DBL detrainment and for regime EW afternoon and nighttime entrainment from AprS (red) and AprN (blue) and from JunN and JunS (black): (a)
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
In Fig. 2a, the turbulent kinetic energy in the detrainment zone between h = hAM and dE(tED) decreases smoothly with both depth and time and by tDS it is nearly uniform at about
6. Empirical turbulence models of the diurnal cycle
The vertical fluxes of buoyancy and momentum are parameterized in terms of the ocean state, surface fluxes and Stokes drift in two distinct ways. The first ETM (K-profile) is based on diffusivity and viscosity in the boundary layer. In the detrainment zone, for example, the rules from section 5a are formulated as analytic profiles of buoyancy flux and stress magnitude, along with
a. K-profile (KP) boundary layer fluxes
These two new parameterizations are evaluated in Fig. 12. Despite differences between hBL and hER (Fig. 7d), the 1:1 lines (dotted) are good fits to cases with Stokes drift (red) and without (blue) over the entire range, though there is somewhat more spread of the black symbols from the WRW period in April and the MRW in June. Nevertheless, all the points of Fig. 12 give correlations of 0.99 for both

Evaluation of the parameterizations of the surface layer TKE production integrals,
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

Evaluation of the parameterizations of the surface layer TKE production integrals,
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Evaluation of the parameterizations of the surface layer TKE production integrals,
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
The K-profile fluxes also depend on a depth scale such as a boundary layer, or entrainment depth. A very simple approach would be to use the MLD as defined by de Boyer Montégut (2004) as the depth where the density (buoyancy) difference from the surface first exceeds a 0.03 kg m−3 threshold. However, this depth remains nearly constant at about 172 m throughout all regimes of AprS. A much smaller threshold is needed to detect the diurnal cycle. For example, with 0.006 kg m−3, the modified

Comparison of (a) nondimensional buoyancy flux distribution from AprS to the parameterizations of (b) ⟨wθ⟩FP (Table 2) and (c) ⟨wθ⟩KP from (26). The contours are irregular at 0.0, ±0.05, ±0.10, ±0.30, and 0.60, with negative contours gray. The time series are boundary layer depth h (solid red) and dE (dotted red) in (a); hPAR (solid) and
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

Comparison of (a) nondimensional buoyancy flux distribution from AprS to the parameterizations of (b) ⟨wθ⟩FP (Table 2) and (c) ⟨wθ⟩KP from (26). The contours are irregular at 0.0, ±0.05, ±0.10, ±0.30, and 0.60, with negative contours gray. The time series are boundary layer depth h (solid red) and dE (dotted red) in (a); hPAR (solid) and
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Comparison of (a) nondimensional buoyancy flux distribution from AprS to the parameterizations of (b) ⟨wθ⟩FP (Table 2) and (c) ⟨wθ⟩KP from (26). The contours are irregular at 0.0, ±0.05, ±0.10, ±0.30, and 0.60, with negative contours gray. The time series are boundary layer depth h (solid red) and dE (dotted red) in (a); hPAR (solid) and
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
The alternative, hPAR, is adapted from sections 4a and 5. Whenever the buoyancy forcing is unstable, the shallowest depth hER is found where a shallower depth
Figures 8 and 10 also show K-profile fluxes (dotted blue) throughout the boundary layer, with υm,s = 0 in (29) convenient for the purpose of revealing major problems above the entrainment depth. Perhaps the most prevalent (bottom panels) is that during entrainment ΩKP = tan−1(⟨wυ⟩KP/⟨wu⟩KP) tracks the near surface veering of the shear, such that
b. F-profile (FP) parameterizations
To utilize hAM and to address detrainment and afternoon entrainment, the appropriate regimes (section 2b) need to be identified without knowledge of the buoyancy flux, especially DBL and EW that are most problematic for K-profile. For this purpose the key factors are the meteorological forcing (Fig. 1); the Langmuir number, La; the Monin–Obukhov depth L; the local maximum stratification at
Summary of the section 6b construction of F-profile fluxes as polynomial segments and angles from the wind, for each diurnal cycle regimes. A single “+” superscript denotes where a segment’s vertical gradient is zero (e.g.,


There also appears to be a high degree of equilibrium, including the entrainment, in regime EC, so it is also relatively straightforward, with nearly linear flux profiles throughout most of the boundary layer. For buoyancy, however, the entrainment point
In regimes EC, EW, and SBL, there is a meaningful boundary layer depth, hPAR, where (29) can be used to give estimates of nonzero fluxes,
The shortcomings of K-profile fluxes indicate that regime DBL is far from equilibrium, such that fluxes are nonlocal in time as well as depth. Hence, they do not reflect a boundary layer depth. Nonetheless, complete F profiles are constructed, as shown in Table 2. First, (22) is integrated from tED to give a deepening dMAX (Fig. 9d), then the detrainment rule (23), as verified in Fig. 11, gives BMAX diminishing with time. Also, an earlier entrainment depth and buoyancy flux at time tED remain a constant endpoint throughout DBL. Curvature in flux profiles may also reflect dependencies on time history. In particular, regime DBL profiles of ⟨wθ⟩FP are cubic functions of depth over three segments that connect with continuous gradients at both dMAX and dE(tED). The deepest segment is cosmetic rather than physical, but smoothly goes to zero at the inversion. Also, τFP is a cubic from
c. Evaluation
For evaluation purposes, υs = 0.1 cm2 s−1 in (29), in rough accord with regime EC. Although primarily for cosmetic purposes, this diffusivity is consistent with purposeful tracer releases in the pycnocline (Ledwell et al. 1993), but is only about one-fifth the regime EW values. In general
Figure 13 shows that a number of features seen in the diurnal cycle of buoyancy flux in the boundary layer are represented by both schemes. Examples include the collapse of the boundary layer at tUS, variations in the vertical extent of the entrainment zone (negative gray contours), as well as the depth and strength of entrainment. One of the more obvious successes is the response to the daytime solar heating between hours 9 and 17, while the buoyancy forcing remains unstable. At all depths in Fig. 13c the primary response is that of Γs from (8) to variations in B0 as the sun rises then sets. This response only affects Fig. 13b directly at
The statistics compiled in Table 3 provide more quantitative evaluations of different regimes and layers, with row numbers provided for reference. They show the F-profile fluxes to have higher fidelity; sometimes significantly. For example (row A.1), in the surface layer of all regimes the ⟨wθ⟩FP regression gives a correlation coefficient greater than 0.99, a slope very near unity and intercept near zero, there is no mean bias, RM = 1.00, and a small root-mean-square difference, δrms = 0.024. Even though surface layer fluxes are strongly constrained by the surface boundary conditions, the corresponding results from ⟨wθ⟩KP (row A.2) show a high bias (RM = 1.12) and nearly a factor of 3 greater δrms; a quantitative reflection of the effects of variable forcing on the flux–profile relationships (25).
Quantitative evaluation of fluxes over various layers in the referenced figures; F profile (FP, middle) and K profile (KP, bottom) against LES (top). The root-mean-square difference is δrms, the ratio of the means RM is the mean bias, and the depth–time correlation coefficient is rz,t. These results are referenced by row number, which are grouped according to the flux; buoyancy (A.1–A.8), downwind momentum (B.1–B.8), and crosswind (C.1–C.8).


There is less fidelity in the more weakly constrained deeper boundary layer. For example, during regime EC, the positive bias above about 70 m in Fig. 13c, leads to RM = 1.24 (row A.4), while there is a low mean bias in Fig. 13b (RM = 0.92) and about half the δrms (row A.3), but the respective correlations fall only slightly to 0.96 and 0.98. Prominent issues in Fig. 13c are the overly negative fluxes in regime SBL and excessive entrainment in regime EW, which are quantified by RM = 1.17, as well as by nearly twice the δrms and the reduced correlation (row A.6), than found for ⟨wθ⟩FP (row A.5), whose regime EW behavior is directly governed by
The full depth of regime DBL is represented only by ⟨wθ⟩FP. Most importantly, these values are maintained at comparable levels to Fig. 13a for some time after the boundary layer collapse at tSU, including at the depth of the maximum (Fig. 9d), despite the negative buoyancy flux nearer the surface (e.g., Fig. 10, hour 34). In particular, they display a similar decay to zero at all depths below the boundary layer by hour 36, as governed by
d. Momentum flux components
Both the magnitude and orientation of the stress vector contribute to errors in parameterized momentum flux components. Clear manifestations of each being dominant are evident in regimes SBL and EW. In particular, Fig. 14c values greater than 1 after hour 42 are due to very large shears in the calculations of τKP from (27) and (28). On the other hand, the tendency for

Comparison of (a) downwind momentum flux −⟨wu⟩τ to parameterized distributions of (b)
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

Comparison of (a) downwind momentum flux −⟨wu⟩τ to parameterized distributions of (b)
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Comparison of (a) downwind momentum flux −⟨wu⟩τ to parameterized distributions of (b)
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

As in Fig. 14, but for the crosswind momentum flux ⟨wυ⟩τ and the parameterized
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1

As in Fig. 14, but for the crosswind momentum flux ⟨wυ⟩τ and the parameterized
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
As in Fig. 14, but for the crosswind momentum flux ⟨wυ⟩τ and the parameterized
Citation: Journal of Physical Oceanography 51, 10; 10.1175/JPO-D-20-0308.1
Figures 14a and 14b both show the downwind momentum flux decreasing monotonically from
Below the surface layer of regime EC the
Momentum fluxes in the detrainment zone above dE(tED) are short lived, and so perhaps negligible for many purposes. Should they become important, however, the qualitative agreement evident between Figs. 14a and 14b and between Figs. 15a and 15b, offers a path forward, with the angle
e. Comparative case evaluation
Evaluations of boundary layer (σ < 1) fluxes, including stress magnitude, τ, from different cases are quantified in Table 4 by the bracketed δrms, and by the correlations in square brackets. For AprS, the higher fidelity of F-profile scheme in Table 3 is reflected in the smaller δrms and higher correlations, compared to the K-profile, for all four fluxes. This relative performance is closely mirrored in the buoyancy flux from the four other cases, but in the momentum flux only from the other two Stokes forced cases, except for the comparatively high D24S
Evaluation of F-profile (FP) and K-profile (KP) fluxes throughout the boundary layer (σ = −z/h < 1) of different cases, as quantified by the root-mean-square difference from the LES, δrms (parentheses), and the depth–time correlation coefficient (square brackets). All regimes are included, except the June MRW period.


The Table 4 F-profile statistics from AprS compared to AprN and from JunS compared to JunN, quantify the improved performance of all four fluxes with Stokes forcing, over cases without. In the idealized, but unphysical absence of Stokes forcing during AprN and JunN, there is much closer alignment of the wind, stress, and shear vectors throughout the boundary layer. Therefore, (8) provides better estimates of ⟨wυ⟩KP = Γυ than it does in the Stokes cases, especially near the surface. This improved fidelity is quantified by the decreases in K-profile δrms relative to the corresponding Stokes cases, in contrast to the increases in the F-profile values. Considering τ for example, a decrease from 0.53 in AprS to 0.38 in AprN is associated with an F-profile increase from 0.19 to 0.38. In general the AprN and JunN momentum fluxes from both schemes are comparable, though the K-profile δrms becomes the smaller for all three quantities in JunN.
7. Discussion and conclusions
The boundary layer depth is an important upper-ocean parameter, because it bounds the extent of large surface forced local and nonlocal turbulent transports through the shape function GC, and it directly scales the mixing coefficients Km,s (7). However, the multiple options in Table 1 can be problematic. Neither the inversion depth, nor a mixed layer depth is a viable proxy, because they do not track the collapse with stable buoyancy forcing, for example. The depth hL19 is not appropriate throughout much of regimes DBL and SBL when it lies beyond the dMKE bound (Fig. 9b), and of regime EW when Fig. 7b shows it is too deep to be consistent with the entrainment rule (17). Similarly, hKPP is overly deep in regime EW (Fig. 7c). In contrast, both the entrainment depth and buoyancy flux are robust diagnostics of LES buoyancy flux profiles, with h = hBL, consistent with (26). However, the parameterizations of these depths are not straightforward, especially when the degree of equilibrium is a factor. The entrainment rule does give a parameterized hER and consistent
In general, the momentum flux is aligned with neither the wind, nor the shear, though at depth the stress vector tends to remain more oriented with the wind than to the veering shear, such there can be a significant across-shear component. A major effect of Stokes forcing is to accentuate the misalignment between the stress and shear vectors, especially near the surface where the Stokes drift and shear are greatest. For example, during regimes SBL and EW of AprS, the angle between the wind and shear ΩW is greater than 45° near 4-m depth. The Hughes et al. (2020) observations of shear in the diurnal warm layer also show ΩW > 45° below 3 m. Thus, the absence of across-shear momentum flux is a possible explanation of why not one of four turbulence parameterizations correlates with their observations. Hence, downgradient eddy viscosity does not appear to be a reliable assumption. Therefore, a fortunate finding is that throughout the boundary layer the angle Ωτ of the stress from the wind is highly correlated with the angle
The development environment for the entrainment and detrainment rules is limited to the Southern Ocean in autumn, but there is a wide range of variable forcing. Thus, the detrainment rules (22) and (23) evaluated in Fig. 11a, include unstable periods of AprS and JunS, as well as the more familiar AprS detrainment below a stable boundary layer. The full range, from extreme wave conditions, including significant swell (L19), to idealized calm seas is involved in the development of the entrainment rule (17). Although the Coriolis parameter is fixed, a variety of rotational interactions, such as cos ΩPAR in (30), is provided by the counterinertial wind rotation of the WRW period, by the subsequent inertial rotation and by the falling wind of D24S. This rule also applies to the very different entrainment of regimes EC and EW, but it should not be regarded as universal, pending further development. In particular, the function F(δρ) in (17) remains to be determined, but the 0.7 coefficient of (32) is consistent with the function decreasing to zero as the stratification at the entrainment depth vanishes. As a preliminary examination of shallower, much stronger stratification and dominant wind forcing, the entrainment rule has been applied to the Watkins and Whitt (2020) simulations of a hurricane over a coastal shelf. Compared to early in AprS and AprN, the stratification at the inversion is stronger by about a factor of 10, the entrainment forcing of (17) is about the same, even though the integrated TKE production to the inversion is about an order of magnitude higher. The entrainment BE is order 100 times greater, and the boundary layer depth is only about one-tenth. Thus, these results would fit well within the spread of Fig. 6, with F(δρ) equal to about 10, at 10 times greater stratification. In the parlance of the dimensional analysis of similarity theory, this finding supports regarding surface layer TKE source integrals as independent variables (L19a), that determine the dependent variables associated with entrainment, at least in regime EC. Alternatively, integrals to the inversion (Fig. 1) would yield consistent results with no change in F(δρ). However, these deeper integrals have yet to be parameterized.
From section 4, the fractions of positive boundary layer TKE production by Eulerian shear, Stokes shear and buoyancy that go to increasing potential energy rather than dissipating, are 0.020, 0.035, and 0.27, respectively. An interpretation is that buoyancy produces TKE at the largest scales and hence is most efficient at increasing potential energy by entrainment. Furthermore, Langmuir turbulence has larger scales and is more efficient than turbulent motions driven by Eulerian shear. Thus, the effects of smaller-scale, near-surface processes, such as wave breaking and solar penetration, may be relatively small, especially for deep boundary layers. In pure convection (
The K-profile fluxes of section 6 are based on Monin–Obukhov similarity theory in the surface layer, including Stokes effects from L19a as updated in section 6, on rules for local and nonlocal (8) boundary layer transport developed in L19b, and on a parameterized boundary layer depth hPAR. They do not yet extend into the detrainment zone below the boundary layer. The parameterizations are empirical and assume that the forcing, fluxes and local shear and stratification are in sufficient equilibrium for eddy diffusivity and viscosity concepts to be applicable. Accordingly, they appear to be most reliable during regimes SBL and EC apart from the WRW and MRW periods, less so for the afternoon entrainment of regime EW, and unreliable during regime DBL and late in D24S. The prospect of improving detrainment is not promising, because there is no signature of a distinct boundary layer depth in any turbulence distribution.
An interpretation of the differences between schemes in the Li et al. (2019) comparison is that realistic forcing has different effects when equilibrium assumptions may not be valid. In contrast, the F profiles are constructed to apply to nonequilibrium boundary layers and detrainment zones by incorporating some time history. Specific examples are hAM (21), dMAX (22), BMAX (23), and
The F profiles are constructed with continuous first derivatives, so that accelerations and buoyancy tendencies are given directly at any depth, rather than as flux differences that strongly depend on gradients. The consequences for numerical stability, adaptive vertical grids and conservation could be substantial. Also, it would be straightforward to solve a heat equation very near surface, so that the evolving temperature would be most appropriate for bulk formula calculations of the surface sensible and latent heat fluxes. This temperature could be combined with an estimate of the cool skin effect (Fairall et al. 1996) to give the sea surface temperature governing the emission of longwave radiation.
Both parameterizations of section 6 have been applied with some success to three flavors of diurnal cycle. The most important is the one April day when solar radiation is sufficient for the surface buoyancy forcing to become stable and the boundary layer to collapse with detrainment below. Diurnal effects are negligible over the first April day and the second June day when the forcing remains unstable and strong, but there are discernible signals when the unstable forcing is weak during regime DBL of the first June day. However, additional cases with much stronger noon heating found in summer or at lower latitudes, with penetrating solar radiation, and with stronger wind are needed for widespread verification. Nonetheless, the prospect of a general mixing scheme applicable to all regimes is supported by the smooth and orderly time–depth flux variations of Figs. 13–15, with past history clearly involved (Figs. 3, 8, and 10), but not always eddy diffusivity and viscosity that rely on equilibrium with local stratification and shear.
Acknowledgments
This work was made possible by support from the U.S. Department of Energy (DOE) under solicitation: DE-FOA-0001036, Climate and Earth System Modeling: SciDAC and Climate Variability and Change, Grant SC-00126005. The 1-h averaged data are available half-hourly from Version 1.0. UCAR/NCAR DASH Repository (https://doi.org/10.5065/87n8-9r86). The other principal investigators—Todd Ringler, Gokhan Danabasoglu, and Matt Long—are gratefully acknowledged, as are the contributions of Alice DuVivier and Justin Small. A special thanks to Dan Whitt for providing results from the Watkins and Whitt (2020) hurricane simulations archived at https://doi.org/10.6084/m9.figshare.12486056.v5 (irene-les-mean-2304 × 84all.nc). The National Center for Atmospheric Research (NCAR) is sponsored by the National Science Foundation. The four SOFS simulations were forced, initialized and evaluated with data sourced from Australia’s Integrated Marine Observing System (IMOS) - IMOS is enabled by the National Collaborative Research Infrastructure Strategy (NCRIS). They utilized resources of the National Energy Research Scientific Computing Center, a DOE Office of Science User Facility supported by the Office of Science of the U.S. Department of Energy under Contract DE-AC02-05CH11231. The idealized simulations utilized high-performance computing on Yellowstone (ark:/85065/d7wd3xhc) provided by NCAR’s Computational and Information Systems Laboratory.
REFERENCES
Ball, F., 1960: Control of the inversion height by surface heating. Quart. J. Roy. Meteor. Soc., 86, 483–494, https://doi.org/10.1002/qj.49708637005.
Burchard, H., and Coauthors, 2008: Observational and numerical modeling methods for quantifying coastal ocean turbulence and mixing. Prog. Oceanogr., 76, 399–442, https://doi.org/10.1016/j.pocean.2007.09.005.
Canuto, V., Y. Cheng, and A. Howard, 2009: Non-local ocean mixing model and a new plume model for deep convection. Ocean Modell., 16, 28–46, https://doi.org/10.1016/j.ocemod.2006.07.003.
Cheng, Y., V. Canuto, and A. Howard, 2002: An improved model for the turbulent PBL. J. Atmos. Sci., 59, 1550–1565, https://doi.org/10.1175/1520-0469(2002)059<1550:AIMFTT>2.0.CO;2.
Craik, A., and S. Leibovich, 1976: A rational model for Langmuir circulations. J. Fluid Mech., 73, 401–426, https://doi.org/10.1017/S0022112076001420.
Danabasoglu, G., W. Large, J. Tribbia, P. Gent, B. Briegleb, and J. McWilliams, 2006: Diurnal coupling in the tropical oceans of CCSM3. J. Climate, 19, 2347–2365, https://doi.org/10.1175/JCLI3739.1.
de Boyer Montégut, C., 2004: Mixed layer depth over the global ocean: An examination of profile data and a profile-based climatology. J. Geophys. Res. Oceans, 109, C12003, https://doi.org/10.1029/2004JC002378.
Deardorff, J., 1972: Theoretical expression for the countergradient vertical heat flux. J. Geophys. Res., 77, 5900–5904, https://doi.org/10.1029/JC077i030p05900.
Fairall, C., E. Bradley, J. Godfrey, G. Wick, and J. Edson, 1996: Cool-skin and warm-layer effects on sea surface temperature. J. Geophys. Res., 101, 1295–1308, https://doi.org/10.1029/95JC03190.
Filipiak, M., C. Merchant, H. Kettle, and P. L. Borgne, 2012: An empirical model for the statistics of sea surface diurnal warming. Ocean Sci., 8, 197–209, https://doi.org/10.5194/os-8-197-2012.
Grant, A., and S. Belcher, 2009: Characteristics of Langmuir turbulence in the ocean mixed layer. J. Phys. Oceanogr., 39, 1871–1887, https://doi.org/10.1175/2009JPO4119.1.
Harcourt, R., 2015: An improved second-moment closure model of Langmuir turbulence. J. Phys. Oceanogr., 45, 84–103, https://doi.org/10.1175/JPO-D-14-0046.1.
Harcourt, R., and E. D’Asaro, 2008: Large-eddy simulation of Langmuir turbulence in pure wind seas. J. Phys. Oceanogr., 38, 1542–1562, https://doi.org/10.1175/2007JPO3842.1.
Högström, U., 1988: Non-dimensional wind and temperature profiles in the atmospheric surface layer. Bound.-Layer Meteor., 42, 55–78, https://doi.org/10.1007/BF00119875.
Hughes, K., J. Moum, and E. Shroyer, 2020: Evolution of the velocity structure in the diurnal warm layer. J. Phys. Oceanogr., 50, 615–631, https://doi.org/10.1175/JPO-D-19-0207.1.
Kantha, L., and C. Clayson, 1994: An improved mixed layer model for geophysical applications. J. Geophys. Res., 99, 25 235–25 266, https://doi.org/10.1029/94JC02257.
Kraus, E., and J. Turner, 1967: A one-dimensional model of the seasonal thermocline, II. The general theory and its consequences. Tellus, 19, 98–106, https://doi.org/10.3402/tellusa.v19i1.9753.
Kukulka, T., A. Plueddeman, and P. Sullivan, 2013: Inhibited upper ocean restratification in nonequilibrium swell conditions. Geophys. Res. Lett., 40, 3672–3676, https://doi.org/10.1002/grl.50708.
Large, W., 1996: An observational and numerical investigation of the climatological heat and salt balances at OWS Papa. J. Climate, 9, 1856–1876, https://doi.org/10.1175/1520-0442(1996)009<1856:AOANIO>2.0.CO;2.
Large, W., and J. Caron, 2015: Diurnal cycling of sea surface temperature, salinity, and current in the CESM coupled climate model. J. Geophys. Res. Oceans, 120, 3711–3729, https://doi.org/10.1002/2014JC010691.
Large, W., J. McWilliams, and S. Doney, 1994: Oceanic vertical mixing: A review and a model with a nonlocal boundary layer parameterization. J. Atmos. Sci., 32, 363–403, https://doi.org/10.1029/94RG01872.
Large, W., E. Patton, A. DuVivier, and P. Sullivan, 2019a: Similarity theory in the surface layer of large-eddy simulations of the Southern Ocean with waves. J. Phys. Oceanogr., 49, 2165–2187, https://doi.org/10.1175/JPO-D-18-0066.1.
Large, W., E. Patton, and P. Sullivan, 2019b: Nonlocal transport and implied viscosity and diffusivity throughout the boundary layer in LES of the Southern Ocean with surface waves. J. Phys. Oceanogr., 49, 2631–2652, https://doi.org/10.1175/JPO-D-18-0202.1.
Large, W., E. Patton, A. DuVivier, and P. Sullivan, 2021: Large eddy simulation of the southern ocean. Version 1.0. UCAR/NCAR DASH Repository, accessed 22 April 2021, https://doi.org/10.5065/87n8-9r86.
Ledwell, J., A. Wilson, and C. Low, 1993: Evidence for slow mixing across the pycnocline from an open-ocean tracer-release experiment. Nature, 364, 701–703, https://doi.org/10.1038/364701a0.
Li, Q., and B. Fox-Kemper, 2017: Assessing the effects of Langmuir turbulence on the entrainment buoyancy flux in the ocean surface boundary layer. J. Phys. Oceanogr., 47, 2863–2886, https://doi.org/10.1175/JPO-D-17-0085.1.
Li, Q., and Coauthors, 2019: Comparing ocean boundary vertical mixing schemes including Langmuir turbulence. J. Adv. Model. Earth Syst., 11, 3545–3592, https://doi.org/10.1029/2019MS001810.
Martin, P., 1985: Simulation of the mixed layer at OWS November and Papa with several models. J. Geophys. Res., 90, 903–916, https://doi.org/10.1029/JC090iC01p00903.
McWilliams, J., P. Sullivan, and C.-H. Moeng, 1997: Langmuir turbulence in the ocean. J. Fluid Mech., 334, 1–30, https://doi.org/10.1017/S0022112096004375.
Mellor, G., and T. Yamada, 1982: Development of a turbulence closure model for geophysical fluid problems. Rev. Geophys., 20, 851–875, https://doi.org/10.1029/RG020i004p00851.
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.
Monin, A., and A. Obukhov, 1954: Basic laws of turbulent mixing in the surface layer of the atmosphere. Tr. Akad. Nauk SSSR Geophiz., 24, 163–187.
Moum, J., A. Perlin, J. Nash, and M. McPhaden, 2013: Seasonal sea surface cooling in the equatorial pacific cold tongue controlled by ocean mixing. Nature, 500, 64–67, https://doi.org/10.1038/nature12363.
Price, J., R. Weller, and R. Pinkel, 1986: Diurnal cycling: Observations and models of the upper ocean’s response to diurnal heating, cooling and wind mixing. J. Geophys. Res., 91, 8411–8427, https://doi.org/10.1029/JC091iC07p08411.
Reichl, B., and R. Hallberg, 2018: A simplified energetics based planetary boundary layer (ePBL) approach for ocean climate simulations. Ocean Modell., 132, 112–129, https://doi.org/10.1016/j.ocemod.2018.10.004.
Reichl, B., and Q. Li, 2019: A parameterization with a constrained potential energy conversion rate of vertical mixing due to Langmuir turbulence. J. Phys. Oceanogr., 49, 2935–2959, https://doi.org/10.1175/JPO-D-18-0258.1.
Schultz, E., S. Josey, and R. Verein, 2012: First air-sea flux mooring measurements in the Southern Ocean. Geophys. Res. Lett., 39, L16606, https://doi.org/10.1029/2012GL052290.
Sullivan, P., J. McWilliams, and C. Moeng, 1994: A subgrid-scale model for large-eddy simulation of planetary boundary-layer flows. Bound.-Layer Meteor., 71, 247–276, https://doi.org/10.1007/BF00713741.
Sullivan, P., L. Romero, J. McWilliams, and W. Melville, 2012: Transient evolution of Langmuir turbulence in ocean boundary layers driven by hurricane winds and waves. J. Phys. Oceanogr., 42, 1959–1980, https://doi.org/10.1175/JPO-D-12-025.1.
Tennekes, H., 1973: A model for the dynamics of the inversion above a convective boundary layer. J. Atmos. Sci., 30, 558–567, https://doi.org/10.1175/1520-0469(1973)030<0558:AMFTDO>2.0.CO;2.
van Roekel, L., B. Fox-Kemper, P. Sullivan, P. 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.
van Roekel, L., and Coauthors, 2018: The KPP boundary layer scheme: Revisiting its formulation and benchmarking one-dimensional ocean simulations relative to LES. J. Adv. Model. Earth Syst., 117, C05001, https://doi.org/10.1029/2018MS001336.
Wang, D., J. McWilliams, and W. Large, 1998: Large eddy simulation of the diurnal cycle of deep equatorial turbulence. J. Phys. Oceanogr., 28, 129–148, https://doi.org/10.1175/1520-0485(1998)028<0129:LESOTD>2.0.CO;2.
Watkins, C., and D. Whitt, 2020: Large-aspect-ratio structures in simlulated ocean surface boundary layer turbulence under a hurricane. J. Phys. Oceanogr., 50, 3561–3584, https://doi.org/10.1175/JPO-D-20-0134.1.
Wyngaard, J. C., 2010: Turbulence in the Atmosphere. Cambridge University Press, 393 pp.