1. Introduction
Sea surface temperature plays an important role in air–sea interaction, and coupled climate models require its accurate prediction. Observations in the Pacific equatorial ocean in recent decades reveal that temperature in the mixed layer (ML) is strongly affected by the subsurface turbulent heat flux associated with deep-cycle turbulence (DCT), which occurs at night on the upper flank of the Equatorial Undercurrent (EUC) (Moum and Caldwell 1985; Hebert et al. 1992; Lien et al. 1996; Moum et al. 2013; Warner and Moum 2019). Recent large-scale regional models (Pei et al. 2020; Cherian et al. 2021) also indicate the importance of the DCT subsurface heat flux in this region, as well as at the Atlantic equatorial ocean where the undercurrent has a diurnal cycle of shear and stratification (Wenegrat and McPhaden 2015) and DCT (Moum et al. 2022).
The DCT is supported by marginal instability (MI) since the gradient Richardson number (Rig) in the layer where it develops clusters around the critical value of 0.25 for shear instability (Smyth and Moum 2013; Smyth 2020). Previous studies suggested that the heat flux depends not only on flow conditions local to the MI layer but also on surface forcing (Lien et al. 1995; Skyllingstad et al. 1999; Smyth et al. 2021, hereafter S21). It also leads to seasonal warming of the ML during boreal spring and cooling during autumn (Moum et al. 2013; Pham et al. 2017).
S21 used moored observations to identify the most important parameters governing the deep cycle, focusing on the parameterization of the turbulent kinetic energy dissipation rate. Using large-eddy simulations (LES), Whitt et al. (2022, hereafter W22), introduced a parameterization for the turbulent buoyancy flux that used different control parameters from those in S21. In the present study, we extend the work of S21 and W22 to examine the seasonal cycle of turbulent heat flux. To do this, we employ LES to characterize DCT over multiple time scales in a parametric study that spans the range of conditions in observations.
On the daily time scale, previous observational and numerical studies suggest the depth-averaged dissipation rate (ε) of the DCT scales with the friction velocity (
Parameterization studies have inherent limitations linked in part to the observational data employed. Shipboard profiling experiments yield detailed vertical resolution but short time series so that surface forcing and background flow condition are limited to a narrow range. The χ-pods yield long time series but relatively coarse vertical resolution which can affect the parameterization quality. The LES simulation in W22 produces high-resolution turbulence data but the forcing parameters are limited to the conditions during the period between October and November 1985. Expanding from the work of W22, we use high-resolution LES to simulate DCT over a wide range of forcing and, additionally, utilize the long-term turbulence measurements from the χ-pods to understand what causes the difference between the scalings in the previous studies and further explore other parameters that can influence the turbulence.
To evaluate the parameterization of DCT, it is necessary to characterize the DCT at turbulence time scales (i.e., minutes to hours). Previous small-scale studies (using shipboard microstructure data or LES) have shown the evolution of the DCT to be complex. It involves the evening descent of a near-surface shear layer which triggers local shear instabilities in the MI layer leading to DCT (Smyth et al. 2011, 2013; Pham et al. 2013). Often, DCT occurs in multiple bursts and exhibits high temporal and spatial fluctuation (M89; Smyth et al. 2017) and so does the heat flux (Sarkar and Pham 2019). Wang and Müller (2002) showed that DCT can be triggered by nighttime convection in the absence of wind stress. Nighttime convection affects the parameterization of turbulence in the surface mixed layer (Lombardo and Gregg 1989); however, its role in the DCT parameterization has not been evaluated. Here, we characterize DCT in a systematic parametric study using LES, specifically addressing the following questions regarding the response of DCT and the associated turbulent heat flux:
-
How would the variation in the wind stress, surface buoyancy fluxes, and background conditions that are found in observations affect the intensity and vertical extent of the bursts of DCT?
-
What importance does convection have in the DCT relative to other forcings such as wind and EUC shear?
-
Can the heat flux be parameterized using the LES data? How does a linear combination of governing parameters compare with previous parameterization? Since the LES is designed to span a wide range of EUC conditions, how does the LES-based parameterization compare with the seasonal cycle in observations?
The paper is organized as follows. The parameter space and description of the LES model are provided in section 2. We address the questions posed above regarding the DCT response at turbulence time scales and the role of convection in section 3. The third question is answered in the affirmative and a new parameterization for daily averaged subsurface heat flux is introduced in section 4. We then apply the parameterization in the context of the seasonal cycle of the heat flux using the long-time data record from the 0°, 140°W TAO mooring in section 5. Discussion and conclusions are given in section 6.
2. Model setup
We construct a parametric study to investigate the characteristics of DCT over a wide range of surface forcing and background flow conditions. The parameters of interest include westward surface wind stress (τw), nighttime convective flux (Qns), shear (S), and the thickness (hMI) of the MI layer. The sign convention for the surface heat flux is positive when downward into the ocean. Here, hMI is calculated as the distance between the MLD and the depth where
Definition of parameters used in discussion.
Values of surfacing cooling flux (Qns), initial thickness of MI layer (hMI), wind stress (τw), and initial shear magnitude in the MI layer (S0) used in parametric studies.
Profiles of zonal current, temperature, squared shear, squared buoyancy frequency, and gradient Richardson number are shown in Figs. 1a and 1b. Since the intensity of the DCT is anticipated to vary with surface heat flux (equivalently surface buoyancy flux B0) and the product of
Initial background profiles used in the parametric study of (a) shear and (b) thickness of MI layer. The panels from left to right show the zonal current (U), temperature difference from a reference value (T − T0), squared shear rate (S2), squared buoyancy frequency (N2), and gradient Richardson number (Rig). (a) Profiles with variable shear in the DCL: S0 = 0.01 (black), 0.015 (red), 0.02 (blue), 0.025 (green) s−1 and (b) profiles with variable thickness of the DCL: hMI = 30 (black), 40 (red), 50 (blue), 60 (green) m. Black dashed lines on the far right panel mark the critical value of 0.25 for linear shear instability. (c) The wide range of the turbulent energetics in the 27 simulations in the
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
In each simulation, there is a spinup period in which the ML responds to the applied wind stress. The duration of spinup varies among the cases depending on surface forcing. We allow the ML to develop until peak entrainment (negative) buoyancy flux reaches 25-m depth. After spinup, we reset time to zero and continue the simulations for 24 h. The surface fluxes, i.e., the wind stress and nighttime buoyancy flux, are turned off after the first 12 h such that the surface mixed layer becomes quiescent during the final 12 h, which corresponds to the daytime period. Previous studies have suggested formation of a diurnal warm layer (DWL) during weak wind conditions (Hughes et al. 2020a,b, 2021). Resolving the DWL requires finer grid resolution near the surface which in turn significantly elevates computational cost. Based on our previous studies (Pham et al. 2017; Sarkar and Pham 2019), the turbulent heat flux driven by the DWL occurs at shallow depths during late afternoons and is distinct from our subject of interest—the nighttime DCT in the MI layer. Therefore, to control the computational cost of the LES suite of 27 simulations, we exclude daytime warming from our examination of DCT.
The LES model is based on the Navier–Stokes equations under the Boussinesq approximation for the evolution of velocity components, dynamical pressure, and temperature. The solver uses second-order finite-difference method for spatial derivatives and mixed third-order Runge–Kutta and Crank–Nicolson methods for time advancement. A multigrid method is used to solve the Poisson equation for the pressure. The parameterization for LES subgrid stresses utilizes the filter structure function as described by Ducros et al. (1996). A subgrid Prandtl number of unity is assumed to compute the subgrid heat fluxes. Further details of the LES implementation can be found in Pham et al. (2023). We assume that stratification is dominated by the temperature gradient, and thus, we neglect the transport equation for salinity [see supplemental information in Smyth and Moum (2013)]. The effects of Langmuir turbulence and precipitation, which were included by Pham et al. (2023), are also neglected. Rain can impact mixing in the ocean surface layer (Smyth et al. 1997; Thompson et al. 2019; Moulin et al. 2021) and particularly at the TAO mooring (Whitt 2022). Like DWL, resolving rain layers in LES greatly increases computational cost, and thus, we do not include effects of rain in the present study.
The computational domain is a rectangular box with size of 384 m × 96 m × 143 m in the zonal (x), meridional (y), and vertical (z) direction, respectively. The grid size of 384 × 96 × 384 points in the three directions gives 1-m resolution in the horizontal direction. Vertical resolution is 0.25 m in the upper 90 m and the grid is mildly stretched at 3% below. Periodicity is imposed in the horizontal directions. Wind stress and buoyancy flux are applied at the top surface. A free-slip condition with constant temperature gradient is applied at the bottom surface. A sponge layer is included in the bottom 20 m to prevent the reflection of internal waves.
To diagnose DCT, we examine the evolution of mean and turbulence statistics. Reynolds averages are denoted by angle brackets (〈⋅〉), which are obtained by computing horizontal averages. Turbulence statistics are obtained as averages of quantities involving fluctuations from the mean, which are denoted by primes. Turbulence quantities of primary interest are the turbulent heat flux (Qt) and the terms in the turbulent kinetic energy (TKE) budget. The heat flux Qt is obtained by computing the sum of resolved and subgrid heat fluxes ρ0cp〈T′w′〉 + Q3,sgs. The major source of the TKE is the shear production SP = −〈u′w′〉∂〈u〉/∂z while the largest sink is dissipation
In the present model setup, the ML and MI layers are idealized regimes where turbulence is energized by surface fluxes and by mean shear, respectively. Although they are never realized precisely in the real ocean, these idealizations provide a useful conceptual picture and a guide for parameterizations. Turbulence in the MI layer is modulated by the surface momentum flux, but it is nevertheless distinct from ML turbulence because its primary energy source is the mean shear in the upper flank of the EUC. This mean shear is not directly set by the local wind and upper-ocean stratification, but rather depends on the equatorial ocean dynamics that occurs at larger spatial and temporal scales. This is the key distinction from the parameterization of turbulence in the mixed layer, in which only the local surface forcing and mixed layer depth are required (Pearson et al. 2015). The ML base that separates the ML and MI regimes is defined in the present study by a temperature drop of 0.04 K from the shallowest value (equivalent to a density drop of 0.01 kg m−3). The MI layer is taken to extend from the MLD to the lowermost depth at which Rig ≤ 0.3. While never exact, these criteria are standard in the interpretation of observations and are used here for consistency with the existing literature.
3. Characteristics of nighttime deep-cycle turbulence
Nighttime DCT occurs in multiple bursts during which the instantaneous dissipation rate (ϵ) varies significantly in time and takes values much larger than inside the mixed layer (M89; Smyth et al. 2013; Pham et al. 2013; W22). Vertical profiles of ϵ show a local peak inside the MI layer below which ϵ decreases smoothly until the capping depth where Rig increases sharply. In this section, we examine how the forcing parameters affect the bursts of DCT with particular attention to the intensities and vertical distributions of ϵ and Qt.
a. Effects of nighttime convection
It is of interest to examine the role of the convective flux in the cases where the wind stress and the shear are weak and B0 is comparable to the product
(a)–(d) Comparison of deep-cycle turbulence among four cases with increasing convection (C1–C4): (left) turbulent dissipation rate, (center) daily averaged profiles of squared shear (S2) and stratification N2, and (right) daily averaged profiles of gradient Richardson number (Rig) and turbulent heat flux (Jq). The nighttime cooling flux takes the values of 100, 125, 150, and 175 W m−2 in the four cases in (a)–(d), respectively. Dashed magenta lines indicate MLD.
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
b. Effects of initial MI-layer thickness
We now consider the impacts of increasing hMI from 30 to 60 m (T1–T4) in the left column of Fig. 3. In cases T1–T3, the second burst persists for a significantly longer time than the first. In contrast, the two bursts in case T4 persist for a similar amount of time. The bursts extend to greater depth in cases with thicker initial hMI. Furthermore, the second burst extends deeper than the first burst and peak values of ϵ during the second burst tends to be stronger. Due to entrainment during the first burst, the instantaneous hMI thickens during the second burst, and the thicker MI layer allows the burst to extend deeper into the thermocline.
As in Fig. 2, but among the four cases with increasing MI layer thickness (T1–T4). The MI layer thickness takes the values of 30, 40, 50, and 60 m in the four cases in (a)–(d), respectively.
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
Peak values of Jq (−84 W m−2 in T2 and −126 W m−2 in T4) tend to increase with increasing hMI although the difference between cases T2 and T3 is minimal. The peak value of daily averaged shear (S) in case T4 is considerably smaller than in the other three cases (see second column of Fig. 3). However, the T4 case has the largest bulk shear computed as an average over the MI layer. This suggests that the peak value of Jq does not increase with the peak value of S in the MI layer, but rather, correlates with the shear at the peak Jq.
c. Effects of wind stress
Increasing the magnitude of wind stress from τw = −0.025 to −0.2 N m−2 (W2–W5) intensifies the DCT. The peak value of ϵ is weakest in the W2 case and strongest in the W5 case (see left column in Fig. 4). More importantly, the turbulence occupies the DCL over a longer time period with increasing wind stress. Only one relatively short burst occurs in the W2 case while there are two bursts in the other three cases (noting that we do not count the small pulse occurring at 18 h in the W5 case as a separate burst). Previous studies have suggested that the first DCT burst of the night is triggered by a descending shear layer from the surface mixed layer (Price et al. 1986; Smyth et al. 2013; Pham et al. 2013, 2017). The descending shear layer was found to form in late afternoon when the solar heat flux relaxes. In the present study, we also observe the formation of a shear layer (not shown) at the MLD prior to the DCT bursts. It deepens into the MI layer and triggers the bursts.
As in Fig. 2, but among the four cases with increasing wind stress (W2–W5). The wind stress magnitude takes the values of 0.025, 0.05, 0.075, and 0.1 N m−2 in the four cases in (a)–(d), respectively, while the shear in the MI layer is held constant at S = 0.02 s−1 (S3).
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
As the wind stress increases from W2 to W5, the first burst occurs at progressively earlier time. In case W2, it takes nearly 9 h for the descending shear layer to penetrate into the DCL and trigger the turbulent burst compared to about 2 h in case W5. Thus, the descent rate of the shear layer is controlled by the wind stress, i.e., the entrainment rate ue of the descending shear layer correlates with the friction velocity
Increasing τw results in stronger daily averaged heat flux (see third column in Fig. 4). The peak value of Jq increases substantially with increasing wind stress. Furthermore, the vertical extent of the Jq profile in case W2 is notably thinner than in the other three cases. When the DCT is relatively weak as in case W2, the turbulent heat flux profile does not spread over the entire vertical extent of the MI layer. In the other three cases, the heat flux diminishes only at depths where Rig exceeds 0.25 below the MI layer.
d. Effects of the MI-layer shear
Increasing shear in the MI layer also results in stronger DCT (see first column in Fig. 5). The peak value of ϵ increases from S1 to S4 cases. In case S1, one long burst of DCT persists throughout the night. In case S2, weak high-frequency modulation is evident. Cases S3 and S4 show two bursts with significant high-frequency modulation. Since the value of Rig in the MI layer is set initially to 0.25 in all cases, the initial buoyancy frequency (N) increases proportionally to the initial shear. Increasing N and S among these four cases promotes growth and modulation of DCT. Due to the increase of DCT with increasing MI-layer shear, the daily averaged MLD becomes shallower (by almost 14 m from S1 to S4). Peak values of Jq are larger with increasing shear (see second and third columns in Fig. 5).
As in Fig. 2, but among the four cases with increasing shear in the MI layer. The shear (S0) takes the values of 0.05, 0.1, 0.15, and 0.2 s−1 in the four cases in (a)–(d), respectively while the wind is held constant at τw = −0.075 N m−2 (W4).
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
The peak values of Jq are similarly large in the high-shear S4 case of this series and the high-wind W5 case of the wind stress variation series. However, there is a difference in the DCT profiles. As noted above, large MI-layer shear tends to spread the DCT turbulence upward and shallows the mixed layer (by about 14 m from S1 to S4). Increasing wind (cf. W5 to W2 in the rightmost panel of Fig. 4) has little effect on the daily average MLD.
While DCT is enhanced when MI-layer shear is increased (i.e., Fig. 5), τw also contributes to increased turbulent mixing (i.e., Fig. 4). The dependence of DCT on both S and τw highlights unique characteristics of forced MI and the resulting DCT. The MI state represents more than just a local (to the EUC) shear instability because the triggering mechanism is as important as the local shear with respect to turbulence energetics. For instance, the entrainment rate of the descending shear layer depends on the wind stress and correlates with DCT and its heat transport as was illustrated in section 3c. Furthermore, the instantaneous dissipation induced by DCT is influenced by the shear and stratification local to the MI layer. The stronger is the shear in the MI layer, the stronger is the DCT. These characteristics of MI and DCT further confirm that the parameterization of ε requires taking into account the strength of surface processes (e.g., wind stress and convection) as well as of the local EUC shear (S21; W22). MI is not unique to the Pacific equatorial ocean. It has also been observed in other oceanic settings such as in Columbia River plume and Mediterranean outflow where the control parameters are different from the ones being considered in the present study (Smyth 2020). Recent laboratory experiments on stratified inclined duct flows (Lefauve et al. 2019) and simulations of forced stratified shear layers (Smith et al. 2021) also show the occurrence of MI in these idealized flow problems. Those studies suggested that the parameterization of the turbulent mixing driven by MI requires a more exhaustive list of control parameters.
e. Nonmonotonic behavior of Jq
While we attempt to extract the DCT dependence on the different forcing parameters in the discussion above, it should be noted that the dependence exhibits variability due to the multiple physical effects involved. This is especially true of the transient behavior of the DCT. The number of turbulent bursts, the peak dissipation in each burst and the persistence of each burst control the net amount of mixing in the DCL over the daily time scale. Figure 6 contrasts Jq in six separate groups. In each, we focus on the correlation between Jq and a particular forcing parameter. Nonmonotonic behavior is observed when the MI layer thickness is varied. The peak value of Jq does not increase between the T2 and T3 cases (i.e., Fig. 6b). In the two groups where we vary τw while holding S in the MI layer constant at S2 and S3 (i.e., Figs. 6c,d, respectively), there is monotonic increase in the peak value of Jq with increasing τw. The monotonic increase is also observed in the group in which S increases while holding τw constant at W3 (i.e., Fig. 6f). However, when we vary S at W3 wind, the peak values of Jq are similar between cases S2 and S3 (i.e., Fig. 6e). The nonmonotonic behavior in the relationship between Jq and the forcing parameters exhibited in some of the groups poses challenges in the parameterization of DCT. Furthermore, since turbulence is a chaotic dynamical system, the bursts of DCT can be sensitive to small perturbations of the background flow and the surface forcing (Liu et al. 2022). Nevertheless, within the parameter space of the present study (which is representative of conditions observed in the Pacific equatorial ocean), we see generally larger Jq in cases with larger τw and S (i.e., Figs. 6c,d,f). Thus, we are able to identify from the LES that τw and S0 have stronger influence on turbulent heat flux than Qns and hMI, similar to the results found in S21 and W22.
Comparison of daily averaged turbulent heat fluxes in various parametric studies: (a) convection, (b) MI layer thickness, (c) wind with moderate shear (S2), (d) wind with strong shear (S3), (e) shear with moderate wind (W3), and (f) shear with strong wind (W4).
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
4. Parameterization of daily averaged deep-cycle turbulence
As the forcing parameters (i.e., Qns, hMI, τw, and S) are varied in control simulations as discussed in the previous section, we observe the following trends in the DCT and the resulting daily averaged heat flux (Jq):
-
The intensity of Jq, especially at its larger values, is mostly influenced by τw and S. Increasing Qns and hMI to high observed values also enhances the DCT but less strongly.
-
While there is a correlation between the MI-layer shear S and the intensity of Jq, peak values of Jq do not occur at the depth with the strongest shear.
-
The intensification of Jq occurs continuously from the surface to the base of the MI layer.
-
Once DCT occurs (e.g., as in Figs. 2–5), the dissipation rate ε tends to increase rapidly relative to the region above the MLD. Thus, the MLD can be used to set the upper bound on the vertical extent of the Jq(z) profile.
Based on these observations, we construct parameterizations for the daily averaged dissipation rate and subsurface turbulent heat flux.
a. Parameterization of ε
To further gain insights to what controls the turbulence in the deep-cycle layer, we analyze the TKE budget at the depth where the turbulent heat flux is most intense (i.e., zmax). Figure 7a shows that the local shear production (Pmax) is the major source of TKE and it is mainly balanced by εmax and the buoyancy flux (
(a) Energetics at zmax indicates the dominant balance between the daily averaged turbulent production (Pmax), dissipation (εmax) and buoyancy flux (
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
The similarity suggests that the shear local to the deep-cycle layer is the main source of TKE. To inspect the role of the surface wind stress, Fig. 7b shows how the momentum flux
(a)–(c) Parameterization of daily averaged dissipation rate. Vertical profiles in (a) are normalized by
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
b. Parameterization of Jq
We aim to parameterize the turbulent heat flux profile, Jq,p(z), specifically the values of heat flux at three depths: the MLD, the base of the MI layer, and the location of peak Jq at z = zmax. These three locations are denoted by the red stars in the schematic diagram in Fig. 9a. Historically, M89 and S21 among other studies of DCT (Lien et al. 1995; Pham et al. 2017) chose to quantify the turbulent mixing in the DCL separately from the turbulence in the mixed layer. Observation 4, discussed at the beginning of this section, also supports the use of MLD as the upper bound of the MI layer. The use of a cutoff gradient Richardson number to set the lower bound of the MI layer (i.e., observation 3 above) is consistent with S21 and W22.
(a) A schematic for a parameterization of the daily averaged turbulent heat flux profile Jq,p(z). The parameterization is constructed using the daily averaged z-dependent profiles of (b) turbulent heat flux (Jq), (c) shear, and (d) gradient Richardson number. The profiles in (b)–(d) only show the values inside the MI layer.
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
Exclusion of the mixed layer from our DCT parameterization has operational advantages as opposed to W22, who chose to include the mixed layer turbulence in their parameterization. The shear data in the surface mixed layer at the TAO mooring is highly limited while the shear in the MI layer is more readily available. By tuning the control parameters (i.e., shear and stratification) to the MI layer instead of the surface mixed layer, we are able to apply the parameterization to the long-term TAO mooring data as will be demonstrated in the next section. For simplicity and ease of comparison with the relatively coarsely-resolved χ-pod data and as previously done in W22, we assume that Jq,p has a piecewise-linear form defined by the values of heat fluxes at these three depths—MLD, zmax, and zMI. With respect to the Jq profile from the LES case shown in Fig. 9a, the linear approximation slightly underestimates the heat flux.
In the rest of this section, we will use the LES suite to find formulae for zmax and zMI and the corresponding values of Jq at these depths as well as at the MLD (the MLD is readily available from the TAO mooring and does not require parameterization). The formulae are deduced based on the daily averaged profiles of Jq, S and Rig from the 27 simulations along with the employed surface forcing (see Figs. 9b–d). We emphasize that the parameterization uses daily averaged background flow conditions from the LES rather than initial conditions in order to account for intraday evolution of DCT in the LES and its effect on the background. There is no loss of generality since the daily averaged shear across the MI layer spans a range as wide as the initial shear (cf. the shear values in Table 2 and the shear profiles in Fig. 9c).
Motivated by observation 3 that Jq decreases to zero at the base of the MI layer, we select the threshold value of Rig = 0.3 to locate zMI and set Jq,p to be zero at this depth. Figure 9b, in conjunction with Fig. 9d to determine the location of Rig = 0.3, shows Jq values at zMI are smaller than 20 W m−2 in nearly all simulations except for the cases with strongest (W6) wind. Using a higher value of Rig for the threshold deepens the MI layer. Such extension in the vertical, although small, deteriorates the agreement between the parameterization and the χ-pod data due to the large shear at zMI (see Fig. 9c). Including the larger shear would be incorrect since it falls outside the region with significant turbulent heat flux. We note that the parameterization in W22 used a slightly larger value of Rig = 0.35 while S21 used Rig = 0.25 for zMI.
Parameterization of (a) the peak heat flux (
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
Parameterization of the depth (zmax) with peak Jq: (a) Jq profiles with marked zmax, (b) shear profiles with marked centroid (zcen), and (c) scaling of zmax and zcen. Panels (a) and (b) only show four cases for illustrative purposes. The solid line in (c) indicates the best linear fit.
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
c. Summary of Jq parameterization
We have developed a parameterization of Jq,p(z) profile intended to infer the turbulent heat flux from basic field observations that do not include microstructure data. The parameterization permits an estimate of Jq, in the upper-ocean surface layer at 0°, 140°W TAO mooring from the available surface forcing, shear, and stratification data since the beginning of the mooring deployment. The Jq,p(z) is a piecewise linear profile that connects the heat fluxes at three locations, signified by the subscript p: zMLD from the observed ML thickness, zMI from the lowest depth where Rig ≤ 0.3 in the observations and, finally, the location, zmax, where Jq peaks, given by Eq. (11). The daily averaged profiles of shear and stratification are sufficient to obtain these three locations. The heat flux is zero at zMI, given by Eq. (9) at zMLD, and given by Eq. (7) at zmax. Equations (7) and (9) require the wind stress and the shear profile as inputs. Thus, with Eqs. (7), (9), and (11), we have deduced a complete formulation to parameterize Jq,p(z) using observed surface forcing and observed shear/stratification profiles.
5. Seasonal cycle of deep-cycle heat fluxes
Previous observational and LES studies have indicated that the DCT exhibits seasonal variability with strong subsurface heat flux during boreal summer months and weaker mixing during the spring and autumn (Moum et al. 2013; Pham et al. 2017; Sarkar and Pham 2019). In this section, we apply the heat flux parameterization (Jq,p) from section 4c to explore its variability in monthly averages. Equations (7), (9), and (11), as summarized in section 4, are used to implement the heat flux parameterization. Using the observed daily averaged profiles of shear and stratification and surface fluxes collected at 0°, 140°W TAO mooring from May 1990 to April 2020, we obtain the parameterized daily averaged heat flux profiles as demonstrated in Fig. 12 for three sample dates. We then bin these daily profiles into the 12 months of the year to deduce the variability of the turbulent heat flux over the seasonal cycle.
Parameterization of daily averaged profiles of turbulent heat fluxes: (a) observed zonal (dashed) and meridional (solid) velocities, (b) observed temperature, and (c) parameterized turbulent heat fluxes. The markers at z = 0 in (c) denote difference between the net downward surface heat flux −Qnet and the penetrative solar heat flux Imax at zmax.
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
Before discussing the seasonal cycle of the turbulent heat flux, we first examine how the surface forcing and background flow conditions change between seasons. Figure 13a shows the observed seasonal variability of solar heat flux Qs, nonsolar heat flux Qns, and τw [see Pham et al. (2017) for further details]. The wind stress shows small values during boreal spring with a minimum in April. The strongest winds occur in August, then slightly decrease during fall increasing again in winter. The strong peaks in August and December are comparable. The ratio
Seasonal variability of (a) wind stress (τw) and solar (Qs) and nonsolar (Qns) surface heat fluxes, (b) the ratio
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
From the daily averaged heat flux profiles (samples are shown in Fig. 12c), we obtain the monthly averaged parameterized peak heat flux
Seasonality of parameterized heat fluxes: (a) peak heat flux (
Citation: Journal of Physical Oceanography 54, 2; 10.1175/JPO-D-23-0015.1
To compare the parameterized heat flux with the observed heat flux from the χ-pods (available at the TAO mooring from September 2005 to March 2019), we compute the bulk (depth average) parameterized heat flux (
In our S21 study, the parameterization was based on dimensional analysis in which the correlation of the observed dissipation rate with
We use daily profiles of Jq,p and Eq. (15) to first obtain the temperature tendency at daily time scale and then average in time to deduce the tendency at seasonal time scale as shown in Fig. 14c. The present parameterization suggests the ocean surface layer is significantly warmed from February to April followed by a period of significant cooling between June and September. There is a net surface layer cooling trend across the annual cycle which agrees with the analysis in Moum et al. (2013).
6. Discussion and conclusions
We have performed a parametric study using LES to identify the processes that influence the DCT in the equatorial Pacific at multiple time scales. DCT is known to occur mostly at nighttime in the marginal instability (MI) layer above the Equatorial Undercurrent (EUC) core. Measurements from the χ-pods deployed at 0°, 140°W between 2005 and 2011 reveal a seasonal cycle of subsurface heat flux driven by DCT that can warm the surface mixed layer (ML) during spring while cooling it during autumn (Moum et al. 2013). We are interested in finding the key physical parameters that affect the heat flux at daily and seasonal time scales. In the LES, we systematically vary surface convective cooling flux (Qns), wind stress (τw), initial thickness (hMI) and shear magnitude (S) of the MI layer over a wide range of observed values. Each case is simulated for a period of 24 h to characterize DCT with high resolution at multiple time scales and to also obtain the daily averaged heat flux (Jq) profile. The daily averaged Jq profiles from the 27 LES cases are analyzed to deduce a new heat flux parameterization that is then demonstrated to provide a reasonable prediction of the seasonal cycle of Jq.
Nighttime DCT exhibits significant variability with multiple bursts of turbulence that persist into the following morning. Among the four control parameters, τw and S are found to most strongly influence daily averaged values of Jq. Increasing τw and S generally increases Jq. Increasing hMI also tends to increase Jq albeit not as strongly as τw and S. Increasing convection slightly enhances the heat flux.
Using the LES results, we formulate a new parameterization for daily averaged turbulent heat flux (Jq,p) that accounts for the effects of the four forcing parameters. This parameterization is summarized in section 4c. Unlike our previous parameterization in S21 and, improving upon that proposed in W22, it uses the local shear within the MI layer along with τw and surface cooling flux to infer the heat fluxes at MLD as well as at the depth zmax where Jq is largest. Heat fluxes at both depths are assumed to be bilinearly proportional to
The present LES-based parameterization builds upon previous studies (i.e., M89; S21; W22; M23) by parameterizing the vertical distribution of the turbulent heat flux and its annual variation in terms of a bilinear combination of buoyancy flux and the product of wind stress and current shear. We apply the new parameterization using TAO mooring data, and the result shows reasonable agreement with χ-pod measurements (see Fig. 14). The peak heat flux values, and the depths at which they occur suggest that strong seasonal cooling during the summer months (June–August) leads to a net annual cooling similar to the observed trend previously reported by Moum et al. (2013). Although the new parameterization includes the effects of convection, its contribution to the seasonal cycle of heat fluxes is usually secondary. The contribution of convection can be larger during periods with weak winds such as the onset of El Niño events (Luther et al. 1983).
Different from M89, S21, and M23, whose parameterizations target the depth-averaged dissipation rate, the new parameterization describes the vertical distribution of the turbulent heat flux as does W22. Although both W22 and the new parameterization propose scaling by the friction velocity and local shear, the new parameterization is calibrated using shear local to the MI layer while the sheared region in W22 extends to include the mixed layer. The parameterization of zmax is also calibrated here using the available flow quantities inside the MI layer, while the zmax in W22 requires information in the surface mixed layer. Since the data at the TAO mooring is limited in the surface mixed layer, the new parameterization improves upon the W22 scheme by circumventing those limitations.
As with any proposed parameterization, it is important to point out its uncertainties. Unlike the LES in W22 and Wang et al. (1998), which includes the effects of large-scale forcing terms, the LES in the present study does not account for the turbulent mixing in the MI layer that can be driven by other processes such as lateral advection. The use of uniform shear and uniform stratification in the initial condition and constant surface fluxes throughout the simulation is an explicit choice made to better understand their physical effect on DCT turbulence and also to reduce the number of control parameters in the scaling analysis. Note that the range of shear, stratification, and surface forcing is sufficient to represent a seasonal cycle in observations. Spatial variability of shear and stratification and temporal variability in the surface forcing can affect the regression coefficients. The new parameterization allows exploration of the spatial and temporal variability of turbulent heat fluxes at the daily time scale and longer. The parameterization is not designed for shorter-than-daily time scales, e.g., it is not a mixing model for prognostic calculations at the time stepping scales used in GCMs.
We have not explored the effect of solar heat fluxes and other turbulent processes in the ML such as Langmuir circulations, breaking waves, and tides, etc. In the limit of low wind speed (U10 ≤ 2 m s−1), a diurnal warm layer (DWL) can form and alter the evolution of turbulence in the ML during daytime (e.g., Hughes et al. 2020a). When the DWL dissipates in the afternoon, the resulting turbulence can affect the descending shear layer and, thus, the instantaneous intensity of the DCT. Breaking surface waves and Langmuir circulations can also enhance ML turbulence (Sullivan et al. 2007; Grant and Belcher 2009; Li and Fox-Kemper 2017) and, by changing the divergence of turbulent momentum transport between the ML and the DCL, modify DCT. Internal waves in the DCL can be excited by Langmuir circulations (Polton et al. 2008) and shear instabilities (Pham et al. 2009; Pham and Sarkar 2010). It remains unknown how these internal waves or internal waves propagating into the sheared zone from deeper water would influence the DCT. Whether these additional phenomena would affect the quality of the present parameterization over the seasonal time scale requires future study.
Acknowledgments.
We are grateful for the support provided by National Foundation Science Grants OCE-1851390 for HTP and SS, and OCE-1851520 for WDS, JNM and SW. JNM also acknowledges the support by National Foundation Science Grants OCE-2048631 and 2049145. Computing support on Gaffney and Koehr was provided by High Performance Computing Modernization Program sponsored by the U.S. Department of Defense.
Data availability statement.
Data are published and a link is provided in Pham (2024). Software necessary to reproduce the figures will be published upon acceptance.
REFERENCES
Cherian, D. A., D. B. Whitt, R. M. Holmes, R.-C. Lien, S. D. Bachman, and W. G. Large, 2021: Off-equatorial deep-cycle turbulence forced by tropical instability waves in the equatorial Pacific. J. Phys. Oceanogr., 51, 1575–1593, https://doi.org/10.1175/JPO-D-20-0229.1.
Ducros, F., P. Comte, and M. Lesieur, 1996: Large-eddy simulation of transition to turbulence in a boundary layer developing spatially over a flat plate. J. Fluid Mech., 326, 1–36, https://doi.org/10.1017/S0022112096008221.
Grant, A. L. M., and S. E. Belcher, 2009: Characteristics of Langmuir turbulence in the ocean mixed layer. J. Phys. Oceanogr., 39, 1871–1887, https://doi.org/10.1175/2009JPO4119.1.
Gregg, M. C., E. A. D’Asaro, J. J. Riley, and E. Kunze, 2018: Mixing efficiency in the ocean. Annu. Rev. Mar. Sci., 10, 443–473, https://doi.org/10.1146/annurev-marine-121916-063643.
Hebert, D., J. N. Moum, C. A. Paulson, and D. R. Caldwell, 1992: Turbulence and internal waves at the equator. Part II: Details of a single event. J. Phys. Oceanogr., 22, 1346–1356, https://doi.org/10.1175/1520-0485(1992)022%3C1346:TAIWAT%3E2.0.CO;2.
Hughes, K. G., J. N. Moum, and E. L. Shroyer, 2020a: 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.
Hughes, K. G., J. N. Moum, and E. L. Shroyer, 2020b: Heat transport through diurnal warm layers. J. Phys. Oceanogr., 50, 2885–2905, https://doi.org/10.1175/JPO-D-20-0079.1.
Hughes, K. G., J. N. Moum, E. L. Shroyer, and W. D. Smyth, 2021: Stratified shear instabilities in diurnal warm layers. J. Phys. Oceanogr., 51, 2583–2598, https://doi.org/10.1175/JPO-D-20-0300.1.
Jacobitz, F. G., S. Sarkar, and C. W. VanAtta, 1997: Direct numerical simulations of the turbulence evolution in a uniformly sheared and stably stratified flow. J. Fluid Mech., 342, 231–261, https://doi.org/10.1017/S0022112097005478.
Large, W. G., and P. R. Gent, 1999: Validation of vertical mixing in an equatorial ocean model using large eddy simulations and observations. J. Phys. Oceanogr., 29, 449–464, https://doi.org/10.1175/1520-0485(1999)029%3C0449:VOVMIA%3E2.0.CO;2.
Large, W. G., J. C. McWilliams, and S. C. Doney, 1994: Oceanic vertical mixing: A review and a model with a nonlocal boundary layer parameterization. Rev. Geophys., 32, 363–403, https://doi.org/10.1029/94RG01872.
Lefauve, A., J. L. Partridge, and P. F. Linden, 2019: Regime transitions and energetics of sustained stratified shear flows. J. Fluid Mech., 875, 657–698, https://doi.org/10.1017/jfm.2019.488.
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.
Lien, R.-C., D. R. Caldwell, M. C. Gregg, and J. N. Moum, 1995: Turbulence variability at the equator in the central Pacific at the beginning of the 1991–1993 El Niño. J. Geophys. Res., 100, 6881–6898, https://doi.org/10.1029/94JC03312.
Lien, R.-C., M. J. McPhaden, and M. C. Gregg, 1996: High-frequency internal waves at 0°, 140°W and their possible relationship to deep-cycle turbulence. J. Phys. Oceanogr., 26, 581–600, https://doi.org/10.1175/1520-0485(1996)026%3C0581:HFIWAA%3E2.0.CO;2.
Liu, C.-L., A. K. Kaminski, and W. D. Smyth, 2022: The butterfly effect and the transition to turbulence in a stratified shear layer. J. Fluid Mech., 953, A43, https://doi.org/10.1017/jfm.2022.985.
Lombardo, C. P., and M. C. Gregg, 1989: Similarity scaling of viscous and thermal dissipation in a convecting surface boundary layer. J. Geophys. Res., 94, 6273–6284, https://doi.org/10.1029/JC094iC05p06273.
Luther, D. S., D. E. Harrison, and R. A. Knox, 1983: Zonal winds in the central equatorial Pacific and El Niño. Science, 222, 327–330, https://doi.org/10.1126/science.222.4621.327.
Moulin, A. J., J. N. Moum, E. L. Shroyer, and M. Hoecker-Martínez, 2021: Freshwater lens fronts propagating as buoyant gravity currents in the equatorial Indian Ocean. J. Geophys. Res. Oceans, 126, e2021JC017186, https://doi.org/10.1029/2021JC017186.
Moum, J. N., and D. R. Caldwell, 1985: Local influences on shear-flow turbulence in the equatorial ocean. Science, 230, 315–316, https://doi.org/10.1126/science.230.4723.315.
Moum, J. N., D. R. Caldwell, and C. A. Paulson, 1989: Mixing in the equatorial surface layer and thermocline. J. Geophys. Res., 94, 2005–2022, https://doi.org/10.1029/JC094iC02p02005.
Moum, J. N., A. Perlin, J. D. Nash, and M. J. 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.
Moum, J. N., and Coauthors, 2022: Deep cycle turbulence in Atlantic and Pacific cold tongues. Geophys. Res. Lett., 49, e2021GL097345, https://doi.org/10.1029/2021GL097345.
Moum, J. N., W. D. Smyth, K. G. Hughes, D. Cherian, S. J. Warner, B. Bourlès, P. Brandt, and M. Dengler, 2023: Wind dependencies of deep cycle turbulence in the equatorial cold tongues. J. Phys. Oceanogr., 53, 1979–1995, https://doi.org/10.1175/JPO-D-22-0203.1.
Osborn, T. R., 1980: Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr., 10, 83–89, https://doi.org/10.1175/1520-0485(1980)010%3C0083:EOTLRO%3E2.0.CO;2.
Paulson, C. A., and J. J. Simpson, 1977: Irradiance measurements in the upper ocean. J. Phys. Oceanogr., 7, 952–956, https://doi.org/10.1175/1520-0485(1977)007%3C0952:IMITUO%3E2.0.CO;2.
Pearson, B. C., A. L. M. Grant, J. A. Polton, and S. E. Belcher, 2015: Langmuir turbulence and surface heating in the ocean surface boundary layer. J. Phys. Oceanogr., 45, 2897–2911, https://doi.org/10.1175/JPO-D-15-0018.1.
Pei, S., T. Shinoda, W. Wang, and R.-C. Lien, 2020: Simulation of deep cycle turbulence by a global ocean general circulation model. Geophys. Res. Lett., 47, e2020GL088384, https://doi.org/10.1029/2020GL088384.
Perlin, A., and J. N. Moum, 2012: Comparison of thermal variance dissipation rates from moored and profiling instruments at the equator. J. Atmos. Oceanic Technol., 29, 1347–1362, https://doi.org/10.1175/JTECH-D-12-00019.1.
Pham, H. T., 2024: Data and software for “Deep-cycle turbulence in the upper Pacific Equatorial Ocean: characterization by LES and heat flux parameterization.” Zenodo, https://doi.org/10.5281/zenodo.10552971.
Pham, H. T., and S. Sarkar, 2010: Internal waves and turbulence in a stable stratified jet. J. Fluid Mech., 648, 297–324, https://doi.org/10.1017/S0022112009993120.
Pham, H. T., S. Sarkar, and K. A. Brucker, 2009: Dynamics of a stratified shear layer above a region of uniform stratification. J. Fluid Mech., 630, 191–223, https://doi.org/10.1017/S0022112009006478.
Pham, H. T., S. Sarkar, and K. B. Winters, 2013: Large-eddy simulation of deep-cycle turbulence in an equatorial undercurrent model. J. Phys. Oceanogr., 43, 2490–2502, https://doi.org/10.1175/JPO-D-13-016.1.
Pham, H. T., W. D. Smyth, S. Sarkar, and J. N. Moum, 2017: Seasonality of deep-cycle turbulence in the eastern equatorial Pacific. J. Phys. Oceanogr., 47, 2189–2209, https://doi.org/10.1175/JPO-D-17-0008.1.
Pham, H. T., S. Sarkar, L. Johnson, B. Fox-Kemper, P. P. Sullivan, and Q. Li, 2023: Multi-scale temporal variability of turbulent mixing during a monsoon intra-seasonal oscillation in the Bay of Bengal: An LES study. J. Geophys. Res. Oceans, 128, e2022JC018959, https://doi.org/10.1029/2022JC018959.
Polton, J. A., J. A. Smith, J. A. MacKinnon, and A. E. Tejada-Martínez, 2008: Rapid generation of high-frequency internal waves beneath a wind and wave forced oceanic surface mixed layer. Geophys. Res. Lett., 35, L13602, https://doi.org/10.1029/2008GL033856.
Price, J. F., R. A. Weller, and R. Pinkel, 1986: Diurnal cycling: Observations and models of the upper ocean response to diurnal heating, cooling, and wind mixing. J. Geophys. Res., 91, 8411–8427, https://doi.org/10.1029/JC091iC07p08411.
Sarkar, S., and H. T. Pham, 2019: Turbulence and thermal structure in the upper ocean: Turbulence-resolving simulations. Flow Turbul. Combust., 103, 985–1009, https://doi.org/10.1007/s10494-019-00065-5.
Simonot, J.-Y., and H. Le Treut, 1986: A climatological field of mean optical properties of the world ocean. J. Geophys. Res., 91, 6642–6646, https://doi.org/10.1029/JC091iC05p06642.
Skyllingstad, E. D., W. D. Smyth, J. N. Moum, and H. Wijesekera, 1999: Upper-ocean turbulence during a westerly wind burst: A comparison of large-eddy simulation results and microstructure measurements. J. Phys. Oceanogr., 29, 5–28, https://doi.org/10.1175/1520-0485(1999)029%3C0005:UOTDAW%3E2.0.CO;2.
Smith, K. M., C. P. Caulfield, and J. R. Taylor, 2021: Turbulence in forced stratified shear. J. Fluid Mech., 910, A42, https://doi.org/10.1017/jfm.2020.994.
Smyth, W. D., 2020: Marginal instability and the efficiency of ocean mixing. J. Phys. Oceanogr., 50, 2141–2150, https://doi.org/10.1175/JPO-D-20-0083.1.
Smyth, W. D., and J. N. Moum, 2013: Marginal instability and deep cycle turbulence in the eastern equatorial Pacific Ocean. Geophys. Res. Lett., 40, 6181–6185, https://doi.org/10.1002/2013GL058403.
Smyth, W. D., P. O. Zavialov, and J. N. Moum, 1997: Decay of turbulence in the upper ocean following sudden isolation from surface forcing. J. Phys. Oceanogr., 27, 810–822, https://doi.org/10.1175/1520-0485(1997)027%3C0810:DOTITU%3E2.0.CO;2.
Smyth, W. D., J. N. Moum, and D. R. Caldwell, 2001: The efficiency of mixing in turbulent patches: Inferences from direct simulations and microstructure observations. J. Phys. Oceanogr., 31, 1969–1992, https://doi.org/10.1175/1520-0485(2001)031%3C1969:TEOMIT%3E2.0.CO;2.
Smyth, W. D., E. D. Skyllingstad, G. B. Crawford, and H. Wijesekera, 2002: Nonlocal fluxes and Stokes drift effects in the K-profile parameterization. Ocean Dyn., 52, 104–115, https://doi.org/10.1007/s10236-002-0012-9.
Smyth, W. D., J. N. Moum, and J. D. Nash, 2011: Narrowband oscillations in the upper equatorial ocean. Part II: Properties of shear instabilities. J. Phys. Oceanogr., 41, 412–428, https://doi.org/10.1175/2010JPO4451.1.
Smyth, W. D., J. N. Moum, L. Li, and S. A. Thorpe, 2013: Diurnal shear instability, the descent of the surface shear layer, and the deep cycle of equatorial turbulence. J. Phys. Oceanogr., 43, 2432–2455, https://doi.org/10.1175/JPO-D-13-089.1.
Smyth, W. D., H. T. Pham, J. N. Moum, and S. Sarkar, 2017: Pulsating turbulence in a marginally unstable stratified shear flow. J. Fluid Mech., 822, 327–341, https://doi.org/10.1017/jfm.2017.283.
Smyth, W. D., S. J. Warner, J. N. Moum, H. T. Pham, and S. Sarkar, 2021: What controls the deep cycle? Proxies for equatorial turbulence. J. Phys. Oceanogr., 51, 2291–2302, https://doi.org/10.1175/JPO-D-20-0236.1.
Sullivan, P. P., J. C. McWilliams, and W. K. Melville, 2007: Surface gravity wave effects in the oceanic boundary layer: Large-eddy simulation with vortex force and stochastic breakers. J. Fluid Mech., 593, 405–452, https://doi.org/10.1017/S002211200700897X.
Thompson, E. J., J. N. Moum, C. W. Fairall, and S. A. Rutledge, 2019: Wind limits on rain layers and diurnal warm layers. J. Geophys. Res. Oceans, 124, 897–924, https://doi.org/10.1029/2018JC014130.
VanDine, A., H. T. Pham, and S. Sarkar, 2021: Turbulent shear layers in a uniformly stratified background: DNS at high Reynolds number. J. Fluid Mech., 916, A42, https://doi.org/10.1017/jfm.2021.212.
Wang, D., and P. Müller, 2002: Effects of equatorial undercurrent shear on upper-ocean mixing and internal waves. J. Phys. Oceanogr., 32, 1041–1057, https://doi.org/10.1175/1520-0485(2002)032%3C1041:EOEUSO%3E2.0.CO;2.
Wang, D., J. C. McWiliams, and W. G. 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%3C0129:LESOTD%3E2.0.CO;2.
Warner, S. J., and J. N. Moum, 2019: Feedback of mixing to ENSO phase change. Geophys. Res. Lett., 46, 13 920–13 927, https://doi.org/10.1029/2019GL085415.
Wenegrat, J. O., and M. J. McPhaden, 2015: Dynamics of the surface layer diurnal cycle in the equatorial Atlantic Ocean (0°, 23°W). J. Geophys. Res. Oceans, 120, 563–581, https://doi.org/10.1002/2014JC010504.
Whitt, D. B., 2022: Rain impacts on the surface atmosphere and upper ocean in the central equatorial Pacific. J. Phys. Oceanogr., 52, 1969–1991, https://doi.org/10.1175/JPO-D-22-0033.1.
Whitt, D. B., D. A. Cherian, R. M. Holmes, S. D. Bachman, R.-C. Lien, W. G. Large, and J. N. Moum, 2022: Simulation and scaling of the turbulent vertical heat transport and deep-cycle turbulence across the equatorial Pacific cold tongue. J. Phys. Oceanogr., 52, 981–1014, https://doi.org/10.1175/JPO-D-21-0153.1.