On the Variation of Dissipation Flux Coefficient in the Upper South China Sea

Jianing Li aCollege of Oceanic and Atmospheric Sciences, Physical Oceanography Laboratory, Institute for Advanced Ocean Study and Frontiers Science Center for Deep Ocean Multispheres and Earth System (FDOMES), Ocean University of China, Qingdao, China

Search for other papers by Jianing Li in
Current site
Google Scholar
PubMed
Close
,
Qingxuan Yang aCollege of Oceanic and Atmospheric Sciences, Physical Oceanography Laboratory, Institute for Advanced Ocean Study and Frontiers Science Center for Deep Ocean Multispheres and Earth System (FDOMES), Ocean University of China, Qingdao, China
bSanya Oceanographic Institution, Ocean University of China, Sanya, China
cLaboratory for Ocean Dynamics and Climate, Pilot National Laboratory for Marine Science and Technology, Qingdao, China

Search for other papers by Qingxuan Yang in
Current site
Google Scholar
PubMed
Close
,
Hui Sun aCollege of Oceanic and Atmospheric Sciences, Physical Oceanography Laboratory, Institute for Advanced Ocean Study and Frontiers Science Center for Deep Ocean Multispheres and Earth System (FDOMES), Ocean University of China, Qingdao, China

Search for other papers by Hui Sun in
Current site
Google Scholar
PubMed
Close
,
Shuwen Zhang dInstitute of Marine Science, Shantou University, Shantou, China

Search for other papers by Shuwen Zhang in
Current site
Google Scholar
PubMed
Close
,
Lingling Xie eLaboratory for Coastal Ocean Variation and Disaster Prediction, College of Ocean and Meteorology, Guangdong Ocean University, Zhanjiang, China
eGuangdong Key Laboratory of Climate, Resources and Environment in Continental Shelf Sea and Deep Ocean, Zhanjiang, China

Search for other papers by Lingling Xie in
Current site
Google Scholar
PubMed
Close
,
Qingye Wang cLaboratory for Ocean Dynamics and Climate, Pilot National Laboratory for Marine Science and Technology, Qingdao, China
gKey Laboratory of Ocean Circulation and Waves, Institute of Oceanology, Chinese Academy of Sciences, Qingdao, China
hUniversity of Chinese Academy of Sciences, Beijing, China

Search for other papers by Qingye Wang in
Current site
Google Scholar
PubMed
Close
,
Wei Zhao aCollege of Oceanic and Atmospheric Sciences, Physical Oceanography Laboratory, Institute for Advanced Ocean Study and Frontiers Science Center for Deep Ocean Multispheres and Earth System (FDOMES), Ocean University of China, Qingdao, China
bSanya Oceanographic Institution, Ocean University of China, Sanya, China
cLaboratory for Ocean Dynamics and Climate, Pilot National Laboratory for Marine Science and Technology, Qingdao, China

Search for other papers by Wei Zhao in
Current site
Google Scholar
PubMed
Close
, and
Jiwei Tian aCollege of Oceanic and Atmospheric Sciences, Physical Oceanography Laboratory, Institute for Advanced Ocean Study and Frontiers Science Center for Deep Ocean Multispheres and Earth System (FDOMES), Ocean University of China, Qingdao, China
bSanya Oceanographic Institution, Ocean University of China, Sanya, China
cLaboratory for Ocean Dynamics and Climate, Pilot National Laboratory for Marine Science and Technology, Qingdao, China

Search for other papers by Jiwei Tian in
Current site
Google Scholar
PubMed
Close
Free access

Abstract

This study focuses on the statistical features of dissipation flux coefficient Γ in the upper South China Sea (SCS). Based on the microscale measurements collected at 158 stations in the upper SCS and derived dissipation rates of turbulent kinetic energy and temperature variance ε and χT, via a modified method, we estimate Γ and analyze its spatiotemporal variation in an energetic and a quiescent region. We show that Γ is highly variable, which scatters over three orders of magnitude from 10−2 to 101 in both regions. Ιn the energetic region, Γ is slightly greater than in the quiescent region; their median values are 0.23 and 0.17, respectively. Vertically, Γ presents a clear increasing tendency with depth in both regions, though the increasing rate is greater in the energetic region than in the quiescent region. In the upper SCS, Γ positively depends on the buoyancy Reynolds number Reb and negatively depends on the ratio of the Ozmidov scale to the Thorpe scale ROT and is scaled as ΓReb1/2ROT4/3, which holds for both regions. The vertical decreasing of ROT is observed, which yields parameterization of ROT = 10−0.002z; this parameterization improves the performance of the Thorpe-scale method by reducing at least 50% of the bias between the observed and parameterized ε. These results shed new light on the spatiotemporal variability and modulating mechanism of Γ in the upper ocean.

Significance Statement

The great global ocean conveyor is maintained by vertical mixing. Turbulent kinetic energy released by local internal wave breaking goes into two parts: one part is used to furnish this vertical mixing, and the rest is dissipated into irreversible heat. The ratio of these two parts is termed as the dissipation flux coefficient and is usually treated as a constant. Our measurements suggest that this coefficient is highly spatiotemporally variable. Specific relationships are obtained when scaling this coefficient with other parameters, and mechanisms modulating this coefficient are also explored. This study sheds light on how much turbulent kinetic energy contributes to elevating the potential energy and its associated influences not only in marginal seas but also in open oceans.

© 2023 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding authors: Q. Yang, yangqx@ouc.edu.cn; H. Sun, sunhui@ouc.edu.cn

Abstract

This study focuses on the statistical features of dissipation flux coefficient Γ in the upper South China Sea (SCS). Based on the microscale measurements collected at 158 stations in the upper SCS and derived dissipation rates of turbulent kinetic energy and temperature variance ε and χT, via a modified method, we estimate Γ and analyze its spatiotemporal variation in an energetic and a quiescent region. We show that Γ is highly variable, which scatters over three orders of magnitude from 10−2 to 101 in both regions. Ιn the energetic region, Γ is slightly greater than in the quiescent region; their median values are 0.23 and 0.17, respectively. Vertically, Γ presents a clear increasing tendency with depth in both regions, though the increasing rate is greater in the energetic region than in the quiescent region. In the upper SCS, Γ positively depends on the buoyancy Reynolds number Reb and negatively depends on the ratio of the Ozmidov scale to the Thorpe scale ROT and is scaled as ΓReb1/2ROT4/3, which holds for both regions. The vertical decreasing of ROT is observed, which yields parameterization of ROT = 10−0.002z; this parameterization improves the performance of the Thorpe-scale method by reducing at least 50% of the bias between the observed and parameterized ε. These results shed new light on the spatiotemporal variability and modulating mechanism of Γ in the upper ocean.

Significance Statement

The great global ocean conveyor is maintained by vertical mixing. Turbulent kinetic energy released by local internal wave breaking goes into two parts: one part is used to furnish this vertical mixing, and the rest is dissipated into irreversible heat. The ratio of these two parts is termed as the dissipation flux coefficient and is usually treated as a constant. Our measurements suggest that this coefficient is highly spatiotemporally variable. Specific relationships are obtained when scaling this coefficient with other parameters, and mechanisms modulating this coefficient are also explored. This study sheds light on how much turbulent kinetic energy contributes to elevating the potential energy and its associated influences not only in marginal seas but also in open oceans.

© 2023 American Meteorological Society. For information regarding reuse of this content and general copyright information, consult the AMS Copyright Policy (www.ametsoc.org/PUBSReuseLicenses).

Corresponding authors: Q. Yang, yangqx@ouc.edu.cn; H. Sun, sunhui@ouc.edu.cn

1. Introduction

Microscale turbulent mixing has always been a research focus, since it furnishes the downward energy cascade by viscously dissipating a part of turbulent kinetic energy (TKE) into irreversible heat (e.g., St. Laurent 2008) and converting the rest of the TKE into background potential energy (e.g., Fernando 1991), which further modulates the large-scale overturning circulation (Wunsch and Ferrari 2004). For a stationary turbulent flow in a finite region, the temporal variation and spatial transport of TKE are often ignored, and the shear production of TKE P, the TKE dissipation rate ε, and the buoyancy flux B due to mixing can reach an equilibrium of P = B + ε. Note that this equilibrium holds only in the case of shear-driven turbulence (Venayagamoorthy and Koseff 2016), because for convection-driven turbulence, the background potential energy becomes a source instead of a sink of TKE (Mater et al. 2015). For the shear-driven turbulence, the partition of TKE converted into buoyancy flux B (to increase background potential energy) defines mixing efficiency, which is actually the flux Richardson number Rif, Rif = B/(B + ε) = B/P, which indicates the ability of turbulence to elevate the background potential energy. The dissipation flux coefficient Γ, Γ = B/ε = Rif/(1 − Rif), is the ratio of the background potential energy gain to the TKE dissipation rate. In some previous literatures, Γ was mistermed as the mixing efficiency. In fact, the value of Rif is always less than unity (Monismith et al. 2018), while that of Γ can be greater than unity in the real ocean (Ijichi and Hibiya 2018).

The dissipation flux coefficient Γ is frequently related to parameterizing diapycnal eddy diffusivity Kρ in ocean general circulation models (St. Laurent et al. 2002) or inferring Kρ from measured ε and stratification N2 (e.g., Lu et al. 2021) based on Osborn’s model Kρ = Γε/N2 (Osborn 1980). In practice, Γ is often taken as a constant of 0.2 (with Rif ∼ 0.17), although there were substantial observations (Moum 1996a; Smyth et al. 2001; Holleman et al. 2016; Ijichi and Hibiya 2018; Monismith et al. 2018; Ijichi et al. 2020; Masunaga et al. 2022), laboratory experiments (Ivey and Imberger 1991; Barry et al. 2001; Jackson and Rehmann 2003), and numerical simulations (Smyth et al. 2001, 2007; Shih et al. 2005; Scotti 2015; Salehipour et al. 2016), which all suggest it varies over several orders of magnitude. Gregg et al. (2018) reviewed the estimates of Γ based on direct measurements of ε and the thermal variance dissipation rate χT in different regions and showed that the value of Γ varies from 10−3 (Monterey shelf; Gregg and Horne 2009) to 1.3 (Admiralty Inlet; Seim and Gregg 1994) and mostly lies in between 0 and 0.3. Recent studies based on microscale observations in various regions indicated Γ scatters in a range from 10−2 to 101 (Mashayek et al. 2017; Ijichi and Hibiya 2018; Ijichi et al. 2020; Vladoiu et al. 2021; Masunaga et al. 2022). Although there are uncertainties induced by estimate bias (Smyth et al. 2001) or the presence of differential diffusion (Jackson and Rehmann 2014), these results all suggest Γ is highly variable.

Apart from statistical analysis, some studies revealed that the variation of Γ is modulated by different dynamic processes and is related to turbulence stages. By conducting direct numerical simulations (DNSs), Smyth et al. (2001, 2007) investigated the evolution of Γ within a whole lifespan of turbulent overturns induced by Kelvin–Helmholtz billows and Holmboe waves, respectively, and found Γ in both cases showed significant time dependence that changed over one order of magnitude (from 0 to 1, approximately). Besides, instabilities triggered by internal waves (Bouruet-Aubertot et al. 2001) and submesoscale motions (Chor et al. 2022) can influence mixing efficiency and regulate the value of Γ. In turn, the variation of Γ modulates the pattern of deep-ocean circulation significantly (Mashayek et al. 2017; Cimoli et al. 2019).

There are studies seeking to reveal the relations between Γ and other variables related to turbulence. Several parameters, such as the density ratio Rρ (Ruddick et al. 1997), the gradient Richardson number Rig (Holleman et al. 2016), and the Froude number Fr (Ivey and Imberger 1991), have been used to interpret the variation of Γ with the hope to reveal the mechanisms controlling Γ and to provide a more reasonable parameterization of Γ. Among those parameters, the most frequently concerned two are the buoyancy Reynolds number (Gibson number) Reb = ε/νN2—where ν is the viscosity coefficient—and the ratio of the Ozmidov scale (Ozmidov 1965) to the Thorpe scale (Thorpe 1977), namely, ROT. Monismith et al. (2018) combined results from DNS, laboratory experiments, and observations and showed Rif remains a constant that is roughly equivalent to 0.2 when Reb < O(100) and is proportional to Reb, namely, RifReb1/2 when Reb > O(100), equivalently indicating Γ remains about 0.25 when Reb < O(100) but decreases with Reb when Reb > O(100). Mashayek et al. (2017) gave an analytical expression indicating a nonmonotonic dependence of Γ on Reb based on several datasets of observations and DNS. Moreover, some studies gave different relations (Salehipour et al. 2016) or even no clear relations (Ijichi and Hibiya 2018) between Γ and Reb. As for ROT, it is often used to indicate the evolution of turbulent patches (Wijesekera and Dillon 1997), with which Smyth et al. (2001) examined the relationship between Γ and ROT and parameterized Γ as Γ0.33ROT0.63. Ijichi and Hibiya (2018) reported a similar dependence of Γ on ROT as ΓROT4/3 and verified it by theoretical scaling. Recent studies also showed a strong negative correlation between Γ and ROT (Ijichi et al. 2020; Vladoiu et al. 2021). As Γ presents strong dependence on ROT, using ROT to parameterize Γ is promising with further verification and improvement (Gregg et al. 2018). Clearly, these works revealed some nature of the variation of Γ; however, Γ cannot be fully parameterized in terms of Reb and/or ROT, as it depends on other variables (Gregg et al. 2018; Monismith et al. 2018).

As the largest marginal sea of the Pacific, the South China Sea (SCS) is known to have elevated turbulent mixing (St. Laurent 2008; Tian et al. 2009) triggered by energetic internal tides (Alford et al. 2011; Klymak et al. 2011; Zhao 2014) and internal lee waves (Buijsman et al. 2012), which presents significant spatiotemporal variation (Yang et al. 2016; Shang et al. 2017; Lu et al. 2021) and deeply modulates the water exchange between the west Pacific and SCS (Qu et al. 2006; Zhou et al. 2014), hence affecting the basin-scale circulation of the SCS. However, few works focused on Γ in the SCS. A few studies (Lu et al. 2014; Sun and Wang 2016) pointed out Γ in the SCS is variable and may be influenced by various dynamic processes, although they only showed limited spatial variations of Γ at several stations, lacking in-depth discussion of the relationships between Γ and turbulence-related variables. Therefore, we examine the spatiotemporal variation features of Γ and explore its relationships to Reb and ROT based on in situ observational data from the upper layer of the SCS. In section 2, we describe the data and methods used in this study. The observed features of Γ and its dependences on Reb and ROT are presented in section 3. Discussion and summary are given in sections 4 and 5, respectively.

2. Data and methods

a. Data

The data used in this study were collected from six cruises in the SCS from April 2004 to April 2011 (Fig. 1). The profiles of microscale velocity fluctuation and fast temperature were obtained by using the Turbulence Ocean Microstructure Acquisition Profiler (TurboMAP; Wolk et al. 2002), which was loosely tethered and nearly free-falling with a speed between 0.5 and 0.7 m s−1 at a sampling rate of 512 Hz (256 Hz at a few stations). Microstructure measurements were obtained at 175 stations, some of which were repeatedly occupied, and a total of 267 casts were available (red squares in Fig. 1). Among these stations, continuous measurements were performed over two days (from 2130 LT 10 May to 0200 LT 13 May 2010) at a station in central SCS with a sampling interval of approximately 1 h (black circle in Fig. 1); 50 casts were obtained at this station after excluding two invalid measurements.

Fig. 1.
Fig. 1.

Station locations (red squares) and bathymetry of the SCS. The black contour marks the isobath of 200 m, which distinguishes the stations used in this study (solid red squares) from those excluded (hollow red squares). The gray box marks the area west of the Luzon Strait, representing region E of the SCS, and the rest of the stations represent region Q. The repeatedly occupied station in the central SCS is marked by a black circle. Topography data are from the ETOPO1 dataset (https://ngdc.noaa.gov/mgg/global/global.html).

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

Microstructure measurements at all stations were limited in the upper layer, with the maximum sampling depth approximately 550 m. At shallow waters, the measurements may be influenced by various dynamic processes near the seabed, and the definition of Γ may not hold near the bottom boundary layer (Mashayek et al. 2017). Therefore, the stations with water depth shallower than 200 m were excluded from data analysis, which included 17 stations and 34 casts (hollow red squares in Fig. 1).

b. Turbulent patch identification

Pioneered by Moum (1996a,b), turbulence-related analyses focus on clear turbulent patches, and many identification techniques were developed. Here, following Mater et al. (2015) and Ijichi and Hibiya (2018), we use the cumulative Thorpe displacement δT to identify turbulent patches, where the Thorpe displacement δT is the depth difference of a water parcel between the original and sorted potential temperature profiles. An example of turbulent patch identification is presented in Figs. 2a–c; the original and sorted potential temperature profiles are coincident at depths of 64.0–65.6 and 73.4–75.0 m, respectively, where the water columns are statically stable, and the corresponding δT and δT remain zero. In contrast, the water column between 65.6 and 73.4 m is significantly overturned, and the values of δT and ΣδT become nonzero and show strong vertical variation. Although temperature, δT, and δT can all indicate the occurrences of overturns, in some complicated cases (e.g., Figs. 2d–f), it is not easy to determine the bounds of a turbulent patch using temperature and δT (Figs. 2d,e). Fortunately, δT remains nonzero within the patch and becomes zero outside the patch (Fig. 2f); as a result, it is adequate to identify turbulent patches.

Fig. 2.
Fig. 2.

Two examples illustrating how the turbulent patch is identified by δT. (a),(d) The vertical profiles of the original and sorted potential temperature θ, (b),(e) δT, and (c),(f) δT.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

The profiles of fast temperature are used to identify turbulent patches. Before identifying turbulent patches, the vertical resolution of the fast-temperature profiles is lowered to 0.1 m to remove fake overturns due to instrument vibration and sensor electronic noise. The measurements in the upper mixed layer are not taken into account, since the falling speed of TurboMAP is not yet stable, and the active mixing events within that layer are mostly convection driven due to the loss of buoyancy flux, and the definition of Γ does not hold anymore. We identify and then exclude the upper mixed layer by employing a threshold method with a relatively large temperature threshold of 0.7°C. This threshold may overestimate the thickness of the upper mixed layer for some stations/casts, but it guarantees the upper mixed layer is thoroughly removed.

To make the results more robust, two criteria are applied to remove some turbulent patches. First, the turbulent patches with vertical size smaller than 1 m are excluded to avoid data distortion. Second, the turbulent patches with −1/2 ≤ Rρ ≤ 2 are excluded to avoid possible contamination by a strongly salinity-stratified layer and density-compensated intrusion (St. Laurent and Schmitt 1999). Here, Rρ=αθz/βSz is the density ratio, where α is the thermal expansion coefficient, β is the saline contraction coefficient, θz and Sz are vertical gradients of potential temperature and salinity, respectively, and the angle bracket denotes an average over each turbulent patch. As Rρ and Turner angle Tu (an indicator of water-column stability to double diffusion) are related by Rρ = −tan(Tu + 45°) (Ruddick 1983), this criterion also avoids the influences of diffusive convection (−90° < Tu < −45°, 0 ≤ Rρ ≤ 1) and strong salt finger (75° < Tu < 90°, 1 ≤ Rρ ≤ 2; You 2002).

As a result, a total of 9641 turbulent patches are identified by examining δT calculated from the fast-temperature profiles. By applying the two criteria, 54.7% of turbulent patches are excluded (corresponding to 5271 patches, among which 5068 were excluded for inappropriate Rρ, 44 for patch size smaller than 1 m, and 159 for both), and 4370 patches remain for analysis.

c. Calculations of Γ, Reb, and ROT

The most widely used method to estimate Γ is based on the assumption that Kρ is equal to the thermal diffusivity KT. Using the fast-temperature measurements, KT is inferred from KT=χT/(2θz2) (Osborn and Cox 1972). By combining Kρ = Γε/N2 and KT=χT/(2θz2) and considering an average over a turbulent patch, an instantaneous patch-averaged Γ is obtained as Γ = 〈χT〉〈N2〉/(2〈ε〉〈θz2) for each turbulent patch. The two parameters, Reb and ROT, are subsequently estimated as Reb = 〈ε〉/(νN2〉) and ROT = LO/LT, where LO = 〈ε1/2/〈N3/2 is the Ozmidov scale and LT=δT21/2 is the Thorpe scale.

The method we use to calculate ε and χT is different from previous studies. To simplify, we take the calculation of ε as an example. Assuming isotropic turbulence and using the Taylor’s frozen assumption, the vertical gradient of velocity uz was derived from a temporal gradient of velocity ut with a falling speed w as uz = ut/w. The vertical wavenumber spectra of uz, φ(k), was then calculated every 2 s with an overlapping of 1 s. In this study, we did not use the often-employed iterative method (e.g., Wolk et al. 2002) to determine the upper integration limit, the Kolmogorov wavenumber kK, and hence ε, but provided a series of Nasmyth spectra φN(k) corresponding to εN = 10e W kg−1 (besides dissipation rate, it requires kinematic viscosity to determine a Nasmyth spectrum, which can be calculated by temperature and salinity; Nasmyth 1970; Oakey 1982; Wolk et al. 2002), where the exponent e varies from −12 to −4 at an interval of 0.1 (gray and blue curves in Fig. 3a). Next, the difference between the observed spectra φ(k) and each Nasmyth spectrum φN(k), Dφ, was integrated in the wavenumber range lower than kK, where kK was determined by the corresponding εN as kK = (εN/ν3)1/4/2π for each φN(k). And ε was finally chosen as εN of the specific φN(k) with the minimum Dφ. The procedure to calculate χT using temporal gradient of fast temperature is similar to that of calculating ε, except that we provided a series of Batchelor spectra ψB(k) corresponding to χTB = 10e °C2 s−1 (Fig. 3b). The upper integration limit of Dψ, the Batchelor wavenumber kB = (ε/νκ2)1/4/2π, is fixed when ε is determined, and κ is the thermal molecular diffusivity.

Fig. 3.
Fig. 3.

Examples of (a),(e) vertical wavenumber spectra of velocity gradient φ(k) (red curve) and of (c) temperature gradient ψ(k) (red curve), together with Nasmyth spectra φN(k) [gray, blue, and yellow curves in (a) and (e); blue has minimum Dφ and yellow is the result estimated by the iterative method] and Batchelor spectra ψB(k) [gray, blue, and yellow curves in (c); blue has minimum Dψ and yellow is the result estimated by the iterative method]. (b),(d),(f)The integrated differences Dφ (or Dψ) between φ(k) and φN(k) [or ψ(k) and ψB(k)] with varying εN (χTB). The gray dots in (a) and (e) mark kK of given εN, whereas kK and kB of observed ε and χT are denoted by the cyan and green lines, respectively.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

As suggested by the examples shown in Fig. 3, the values of Dφ and Dψ are sensitive to the variations of εN and χTB (Figs. 3b,d): they reach their minima when φ(k) and ψ(k) are well fitted by specific φN(k) and ψB(k), respectively, suggesting ε = 2.0 × 10−9 W kg−1 and χT = 1.0 × 10−8 °C2 s−1 in this case (Figs. 3a,c); when εN and χTB deviate from these values, Dφ and Dψ grow exponentially to show the increasing bias between the observed ε (χT) and given εN (χTB). This new method, namely, the difference integration method, gives more accurate estimates than the iterative method when φ(k) cannot be well fitted by φN(k). As shown in Fig. 3e, φ(k) in a wavenumber range lower than 7 cpm is significantly inconsistent with any given φN(k). The iterative method is highly impacted by the uplift of φ(k) in the low-wavenumber range, resulting in an overestimate of ε as 7.2 × 10−9 W kg−1. While the difference integration method takes all wavenumbers smaller than kK into account and gives a smaller estimate as 1.3 × 10−9 W kg−1, whose spectrum fits much better with φ(k) in the wavenumber range between 10 cpm and kK, 33 cpm; Dφ of difference integration method (31) is also smaller than that of iterative method (67), suggesting result of the new method is closer to the truth. Considering the calculation procedure is performed twice for ε and χT, and the estimate of χT is on the basis of the calculation of ε, using the difference integration method could significantly improve the accuracy of the results of “anomalous” spectra. A detailed comparison between these two methods and their influences on the results can be found in the appendix.

The terms N2 and θz are also estimated by the fast-temperature measurements after the resolution is lowered to 0.1 m. As demonstrated in Smyth et al. (2001) and reaffirmed in Ijichi and Hibiya (2018), 〈θz〉 should be examined in different ways for proper estimates of Γ and LO (Figs. 2a,d): the average one 〈θzmean inferred from the linear fitting of the sorted potential temperature should be used to calculate Γ, while the bulk one 〈θzbulk = 〈θ21/2/LT, where θ′ is the difference between the original and sorted potential temperatures, should be used to calculate LO. For turbulent patches consisting of a single, large overturn, as in the example shown in Fig. 2a, 〈θzmean and 〈θzbulk are almost identical, being 0.040 and 0.041°C m−1, respectively. However, when a turbulent patch contains several overturns separated by thin stable layers (Fig. 2d), those thin stable layers are included to calculate 〈θzmean but are largely ignored to obtain 〈θzbulk (θ′ is much smaller), thus 〈θzmean turns out to be greater than 〈θzbulk. In the case presented in Fig. 2d, 〈θzmean is twice as large as 〈θzbulk, being 4.6 × 10−3 and 2.3 × 10−3 °C m−1, respectively.

3. Results

a. Horizontal variation of Γ

Near the Luzon Strait, a large number of vigorous internal tides are generated (Klymak et al. 2011), and those of higher modes break and dissipate locally to furnish elevated mixing (Niwa and Hibiya 2004; Tian et al. 2009; Yang et al. 2016; Shang et al. 2017; Lu et al. 2021). The low-mode internal tides continue to propagate southwestward and enter the SCS, becoming very weak when arriving at the southwestern SCS (Zhao 2014), accompanied by low-level turbulent mixing (Yang et al. 2016; Lu et al. 2021). In this view, we divided the SCS into two regions, the area west of the Luzon Strait (114°–121°E, 17°–23°N), featured with energetic turbulent mixing (region E; marked by a gray box in Fig. 1), and the rest of the SCS deeper than 200 m as the quiescent region (region Q). We explore the features of Γ in these two regions based on 1227 turbulent patches identified from 74 casts in region E and 3143 patches from 145 casts in region Q.

In both regions Γ varies over several orders of magnitude (Fig. 4a); Γ scatters in a larger range in region Q than in region E, with the former from 1 × 10−3 to 15.0 and the latter from 1 × 10−2 to 6.9. This observed variation range of Γ between 10−3 and 101 is similar to various observations conducted in different regions (e.g., Smyth et al. 2001; Ijichi and Hibiya 2018; Ijichi et al. 2020; Vladoiu et al. 2021; Masunaga et al. 2022). The median (mean) value of Γ is 0.23 (0.39) in region E, larger than in region Q, 0.17 (0.30). The distributions of Γ in both regions are slightly positively skewed with skewness of 0.11 in region E and 0.09 in region Q. In region E, Γ follows a nearly mesokurtic distribution with excess kurtosis of 0.09, while that in region Q is a little leptokurtic, with excess kurtosis of 0.71. The maximum, median, and mean values of ε in region E (2.7 × 10−7, 4.8 × 10−10, and 3.4 × 10−9 W kg−1, respectively) are all greater than those in region Q (1 × 10−7, 8.1 × 10−11, and 7.9 × 10−10 W kg−1, respectively; Fig. 4b). The probability distribution of ε in region Q is clearly positively skewed (skewness 1.05) and leptokurtic (excess kurtosis 0.66), and it presents a distinct peak at 4.6 × 10−11 W kg−1, while that in region E is mildly positively skewed (skewness 0.39) and platykurtic (excess kurtosis −0.36), with high probability of ε ranging from 4.2 × 10−11 to 2.5 × 10−9 W kg−1, indicating region E experiences a stronger dissipation than region Q. The difference of χT between the two regions is similar to that of ε (Fig. 4c), and the maximum, median, and mean values of χT are all greater in region E (7.0 × 10−7, 1.4 × 10−9, and 9.0 × 10−9 °C2 s−1, respectively) than those in region Q (5.1 × 10−7, 2.0 × 10−10, and 2.0 × 10−9 °C2 s−1, respectively). The distributions of χT in regions E and Q are both platykurtic (excess kurtosis −0.35 and −0.34, respectively), whereas the latter is more positively skewed than the former, with skewness of 0.57 and 0.13, respectively.

Fig. 4.
Fig. 4.

Histograms normalized by probability of (a) Γ, (b) ε, (c) χT, (d) N2, (e) Reb, and (f) ROT in regions E (red curve) and Q (blue curve). The dots and triangles mark the median and mean values of the corresponding curve, respectively. The gray dashed line in (a) denotes Γ = 0.2.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

In contrast to the distinct features of ε and χT between regions E and Q, the stratification distributions in both regions are similar (Fig. 4d), and their median and mean values are nearly identical, 3.3 × 10−5 and 4.7 × 10−5 s−2, respectively, in region E and 3.1 × 10−5 and 4.6 × 10−5 s−2, respectively, in region Q, although showing slight differences in probability distribution. The probability distribution of N2 in region E is slightly negatively skewed and platykurtic (skewness −0.18, excess kurtosis −0.16), whereas that in region Q is positively skewed and leptokurtic (skewness 0.20, excess kurtosis 0.13). Vladoiu et al. (2021) found that energetic and quiescent regions in the western Mediterranean Basin are contrasted in both dissipation and stratification, while our results suggest the energetic and quiescent regions in the upper SCS are distinguished by distinct dissipation with similar stratification. The probability distributions of Reb and ROT in region E are distinct from those in region Q (Figs. 4e,f), as they are directly derived from distinct ε and similar N2. In region Q, Reb scatters in a range from 1.6 × 10−1 to 3.6 × 103, with median and mean values of 4.0 and 31.8, respectively, which are all smaller than those in region E (from 3.1 × 10−1 to 6.7 × 104, and 19.4 and 234.6). The distributions of Reb in both regions are positively skewed, with skewness of 0.38 in region E and 0.83 in region Q; the excess kurtoses are −0.18 and 0.49, respectively, suggesting the former is platykurtic while the latter is leptokurtic. The scatters of ROT are from 2.7 × 10−3 to 5.8, with median and mean values of 0.21 and 0.35 in region Q versus from 9.7 × 10−3 to 11.0, with 0.42 and 0.69 in region E. And their skewnesses (excess kurtoses) are 0.09 and −0.04 (1.63 and 0.30), respectively. The variation ranges of Reb and ROT in the upper SCS are generally consistent with those found in previous observations (Smyth et al. 2001; Ijichi and Hibiya 2018; Ijichi et al. 2020; Lu et al. 2021; Vladoiu et al. 2021; Masunaga et al. 2022), except that some small values (Reb < 1 and ROT < 0.1) are captured in our study. All these differences between regions E and Q validate that turbulent mixing in region E is more energetic and efficient than that in region Q.

b. Vertical variation of Γ

Ijichi and Hibiya (2018) observed a clear tendency of Γ increasing from surface to 3000 m in the western and central North Pacific Ocean and the Southern Oceans. Smyth (2020) suggested Γ should be around 0.2 away from the upper and lower boundaries, mingled with high-efficiency coherent structures. The variation of Γ with depth is related to the prevalence of coherent structure, which encourages us to explore the vertical variation of Γ in the upper SCS.

The vertical variations of N2 in regions E and Q are characterized by three layers (Fig. 5): the thermocline with strongest stratification N2 > 1.6 × 10−4 s−2 (black contours), a moderate layer with stratification 1.6 × 10−4 s−2N2 ≥ 4.0 × 10−5 s−2 (white contours), and the deeper less-stratified layer with N2 < 4.0 × 10−5 s−2. The thermocline can be observed in all casts in region E, while it is shallow and weak in some casts in region Q. In general, the thermocline depth varies between 100 and 200 m, and the moderate layer depth varies between 200 and 400 m; as a result, we use 150 and 300 m to separate the observed water column into three layers.

Fig. 5.
Fig. 5.

Vertical variation of N2 (s−2) within the upper 500 m in region (a) E and (b) Q. The black and white contours denote the N2 values of 1.6 × 10−4 and 4 × 10−5 s−2, respectively.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

Figure 6 shows the vertical variations of Γ in regions E and Q, along with other turbulence-related variables. The median values of Γ, binned by depth, and their linear fitting present a clear increasing trend with increasing depth in both regions (Fig. 6a). The median values of Γ at 75 m are 0.1 for both regions, and they increase to 0.25 and 0.48 at 500 m in regions Q and E, respectively. The increasing rate of Γ with depth in region E is 2 times higher than that in region Q, being 1.4 × 10−3 and 8.2 × 10−4 m−1 (in logarithmic space), respectively. There are some detailed vertical structures of Γ in different layers: in region Q, Γ increases in a slightly weaker rate within the moderate layer than that in the thermocline and less-stratified layer; in region E, Γ keeps increasing in the thermocline and the moderate layer but shows a somewhat decreasing trend in the less-stratified layer, such as at depth deeper than 450 m. The variation range of Γ (represented by the 5th and 95th percentiles) also shifts to larger magnitude with increasing depth. Within the thermocline, Γ in region Q varies over a wider range than that in region E, with the 5th percentile being smaller and the 95th percentile being larger in region Q than in region E. Beneath the thermocline, Γ in region E varies in a range of larger magnitude than that in region Q, as suggested by the 5th and 95th percentiles in region E being both lager than those in region Q. In general, turbulent mixing tends to become more efficient with increasing depth in both regions; however, in the water column deeper than 150 m, it has a stronger ability to elevate the potential energy of seawater in region E than in region Q. This increasing trend of Γ with increasing depth is similar to the result of Ijichi and Hibiya (2018); Ijichi et al. (2020) also suggested young instabilities with higher Γ are more common at depth due to the weaker stratification. Moreover, Smyth (2020) revealed that highly efficient, long-lasting coherent structures are responsible for intermittent higher Γ in the interior.

Fig. 6.
Fig. 6.

Vertical variations of (a) Γ, (b) ε, (c) χT, (d) N2, (e) Reb, and (f) ROT of each turbulent patch in regions E (small magenta dots) and Q (small cyan dots). The median values binned by depth and their associated linear fittings are marked by big dots and lines, respectively, with red representing region E and blue representing region Q. The 5th and 95th percentiles are marked by dash–dotted lines to outline the variation ranges. The horizontal gray dashed lines in each panel mark the bottom depths of the thermocline and the moderate layer, 150 and 300 m, respectively, and the vertical gray dashed line in (a) denotes Γ = 0.2. The black dots denote those patches with ROT smaller than 0.05 and depth shallower than 150 m. The turbulent patch numbers of each depth bin are presented in (e) for region Q and (f) for region E.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

The terms ε and χT share similar vertical variations (Figs. 6b,c). In general, they decrease with depth in both regions. In region Q, ε and χT weaken sharply within the thermocline and the moderate layer, with ε varying from O(10−9) at 50 m to O(10−11) W kg−1 at 220 m and χT varying from O(10−9) at 50 m to O(10−10) °C2 s−1 at 260 m, respectively. It contrasts sharply with the scenario in the less-stratified layer in region Q, where ε and χT vary slightly and have the same order; ε is basically identical to 7 × 10−11 W kg−1, and χT is mostly O(10−10) °C2 s−1. The vertical variability of ε and χT in region E is similar to that in region Q, although the magnitude is 3–5 times larger in region E than in region Q at different depths. The variation range of ε in region Q becomes narrower with increasing depth, whereas that in region E shows no clear change.

The stratifications in both regions decrease logarithmically with depth from O(10−4) to O(10−5) s−2 (Fig. 6d), presenting no substantial difference. The vertical variation of median values of Reb suggests it generally decreases with increasing depth (Fig. 6e), which varies from 13.3 to 3.6 between 70 and 500 m in region Q and from 27.5 to 16.6 in region E. In region Q, the decrease of Reb is more significant in the thermocline, which varies from 13.3 to 4.0, and it remains at that value beneath the thermocline. Note that the slight decreasing tendency of Reb with increasing depth here is due to the relatively faster decreasing of ε (about two orders in magnitude from 50 to 550 m) compared with the relatively slower decreasing of N2 (about one order of magnitude; Figs. 6b,d). However, the decreasing of Reb in the upper 550 m is opposite to the general variation of Reb when considering the full water depth, where Reb increases several orders of magnitude with increasing depth (Mashayek et al. 2017; Ijichi et al. 2020). The larger Reb in region E than in region Q clearly suggests a higher-level turbulence intensity near the Luzon Strait.

The median values of ROT in region Q (region E) show a consistent decreasing tendency between 70 and 500 m from 0.37 to 0.13 (from 0.72 to 0.22; Fig. 6f). Beneath the thermocline, ROT in region Q varies in a narrower range than that in region E, both of which have lower limit of about 10−1. However, the lower bounds of variation range in the thermocline in both regions bend obviously toward smaller ROT as a result of the intensive occurrences of turbulent patches with ROT smaller than 0.05, which are very rare beneath the thermocline. These extremely small ROT may be related to the life-span of turbulence and are discussed in section 3c in detail. In summary, there is a clear decreasing variability for ε, χT, N2, Reb, and ROT with increasing depth, although at different rates; while Γ increases with depth at a greater increasing rate in region E compared with that in region Q.

c. Variation of Γ related to the life-span of turbulence

We found 75 turbulent patches with ROT smaller than 0.05 within the thermocline in both regions (black dots in Fig. 6; green dots with black edges in Fig. 7), which have the smallest magnitude of ROT compared with the other turbulent patches at all depths. As the Kelvin–Helmholtz instability is believed to be the main mechanism of turbulence generation in the stratified, sheared ocean (Smyth et al. 2001; Mashayek et al. 2017) and ROT can be used to indicate the life-span of such turbulent patch (Smyth et al. 2001; Mater et al. 2015), we infer these patches are in an “initial” stage of turbulence evolvement. We noticed that turbulent patches with extremely small ROT were also captured in Mater et al. (2015), which were driven by different mechanisms away from the thermocline. They demonstrated that ROT can be used as an indicator of turbulence age not only for Kelvin–Helmholtz instability but also for other large-scale instabilities such as convectively driven turbulence.

Fig. 7.
Fig. 7.

Differences of Γ and ε between the young (ROT < 0.5; green dots) and mature stage (ROT > 0.5; yellow dots) of turbulent patches in regions (a) E and (b) Q. The green and yellow squares with black edges are the median values of Γ and ε from the patches in the young and mature stage, respectively, and the dots with black edge are those patches with ROT < 0.05 and depth shallower than 150 m.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

The definition of an initial stage of turbulence life-span is put forward on the basis of the “young” and “mature” stages (Smyth et al. 2001; Mashayek et al. 2017); therefore, it is necessary to verify if the turbulence behaviors in young and mature stages are valid for our observations. Turbulent patches with ROT < 0.5 are regarded as the young stage, whereas patches with ROT > 0.5 are considered as the mature stage (Smyth et al. 2001). Although this criterion is based on numerical simulations with a low Reynolds number, in which the evolution of turbulence might differ from that in the real ocean, we tried to use it to check if the young and mature turbulences can be distinguished in our field observations. As shown in Fig. 7, the observed turbulent patches with ROT < 0.5 are typically distinguished from those with ROT > 0.5 by low ε and high Γ. In region E, the median values of ε and Γ from the turbulent patches with ROT < 0.5 and ROT > 0.5 are (1.5 × 10−10 W kg−1, 0.30) and (1.9 × 10−9 W kg−1, 0.16), respectively, while those in region Q are (6.0 × 10−11 W kg−1, 0.18) and (1.2 × 10−9 W kg−1, 0.12), respectively. These differences are consistent with previous conclusions on the young and mature stages of turbulence, that is, the young-stage turbulence is featured with high efficiency and low dissipation because it represents the two-dimensional Kelvin–Helmholtz billow that overturns the density interfaces (high efficiency) before the turbulent, three-dimensional secondary instabilities are fully developed (low dissipation). This consistency suggests young and mature stages of turbulence can be well identified in our observations.

The turbulent patches with extremely small ROT in the thermocline are highlighted by black edges in Fig. 7. Compared with the other turbulent patches in the young stage, they present significantly higher mixing efficiency, with Γ ≈ 1 in region E and Γ ranging from 0.32 to 6.85 in region Q, which are remarkably greater than the median values of young-stage turbulent patches of 0.30 and 0.18, respectively. As for ε, its median values (1.7 × 10−9 W kg−1 in region E and 5.7 × 10−10 W kg−1 in region Q) are higher than those of young-stage turbulent patches [O(10−10) W kg−1]. However, if examined in the same depth range, the ε values of those patches are smaller than the median values of ε (2.5 × 10−9 W kg−1 in region E and 1.3 × 10−9 W kg−1 in region Q; Fig. 6b). That is to say, these patches are featured with smaller ROT, smaller ε, and greater Γ; therefore, they are definitely “younger” than those turbulent patches in the young stage and hence are defined as initial stage. The initial stage refers to the very beginning stage of turbulence, when the relatively large Kelvin–Helmholtz billow is just formed, and upcoming smaller secondary instabilities are yet to be developed. This is confirmed by the temperature examples shown in Figs. 8a–d. The unsorted temperature profiles suggest the occurrence of overturns. Compared with those of young and mature turbulent patches (e.g., Fig. 2d), these profiles are significantly uncomplicated, containing only several eddies with relatively large vertical scales (represented by “S” structures). Although these profiles are obtained at different stations (Fig. 8e) and dates, the initial Kelvin–Helmholtz billow shed more eddies (from 1 to 3) with ROT increasing from 0.009 to 0.026 (Figs. 8a–d). These structure evolvements, together with the increasing ROT, are consistent with the beginning stage of the evolution of Kelvin–Helmholtz billows (Fig. 1 of Smyth et al. 2001).

Fig. 8.
Fig. 8.

(a)–(d) Examples of temperature profile of the turbulent patches in the initial stage, and (e) the station locations where these patches are observed. The black curves in (a)–(d) are the sorted temperature profiles, whereas the colorful ones are the original temperature profiles; the station locations of these examples are marked by corresponding colors in (e). In (e), the number to the left of the station indicates the number of initial-stage patches at that station.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

It seems that these initial-stage turbulent patches have a unique spatial distribution. Vertically, they are observed within the thermocline, suggesting their energy source is potentially the near-inertial energy input from the upper mixed layer induced by surface wind. Besides, among the total 75 initial-stage patches, only 7 are observed in region E, indicating the initial stage is more easily captured in region Q than in region E. A possible explanation for this regional difference is that the turbulence in region E may require much shorter time to experience its whole lifespan compared with that in region Q, because very energetic internal waves in region E promote turbulence evolving much faster, leading to a relatively shorter transient initial stage and hence is not frequently observed.

d. Dependences of Γ on ROT and Reb

Because more turbulent patches were detected in region Q and hence can give a more representative and solid conclusion, we demonstrate the relationships of Γ with Reb and ROT in region Q in detail (Fig. 9). The results in region E are also given in Fig. 10, showing a similar pattern, except for some quantitative discrepancies.

Fig. 9.
Fig. 9.

Variation of Γ (a),(b) with Reb and (c),(d) ROT in region Q. The general variations of Γ with Reb and ROT are examined by considering all turbulent patches as a whole in (a) and (c), although they are colored by ROT and Reb, respectively. The black dots in (a) and (c) are the median values of Γ binned by Reb (ROT), and the black vertical bars represent the corresponding 10th–90th percentile ranges. The red curves in (a) and (c) represent the parameterizations proposed by Mashayek et al. (2017) and Ijichi and Hibiya (2018), respectively. The solid black line in (c) is the linear fitting of the median values of Γ binned by ROT, while the dashed lines are linear fittings in different ROT ranges. The variations of Γ with Reb and ROT within specific ranges of ROT and Reb are presented in (b) and (d), respectively. The dots with black edges and overlying lines denote the median of Γ binned by Reb or ROT and the corresponding linear fittings, respectively, with colors indicating different ranges of ROT and Reb; fitting presenting slope distinct from 1/2 or −4/3 is marked by dashed line. The red line in (d) represents the parameterization proposed by Ijichi and Hibiya (2018). The gray line in each panel marks Γ = 0.2.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

Fig. 10.
Fig. 10.

As in Fig.9, but for region E.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

The general variations of Γ with Reb and ROT are evaluated separately (Figs. 9a,c). The median values of Γ vary in a range of 0.13–0.25, with Reb increasing from 0.5 to 400, which can almost be taken as a constant when compared with Γ spanning over several orders of magnitude (Fig. 9a). When Reb is greater than 400, Γ drops to about 0.1, which may suggest the decrease of Γ for very strong turbulence intensity, as reported in previous studies (Mashayek et al. 2017; Monismith et al. 2018; Vladoiu et al. 2021), although the insufficient sample size brings bias to this result. This decrease of Γ is not seen in region E, even when Reb exceeds 1000 (Fig. 10a). There is a nonmonotonic relationship between Γ and Reb. The variation of Γ within the range of Reb between 1 and 100 fits especially well with the parameterization formula Γ(Reb)=2Γ*(Reb/Reb*)1/2/[1+(Reb/Reb*)] proposed by Mashayek et al. (2017), where Γ* is the maximum of Γ at Reb=Reb*. The fitting does not work well when Reb < 1 and Reb > 100, which may be due to sample size being inadequate in these ranges. Ιn region Q, Γ* and Reb* are 0.22 and 16, and Γ* = 0.22 is comparable to the lower bound of the datasets reported in Mashayek et al. (2017), while Reb* = 16 is much smaller than O(102) in Mashayek et al. (2017). The median of Γ decreases by an order of magnitude with increasing ROT (Fig. 9c). This negative dependence of Γ on ROT has been reported in DNS and field observations (e.g., Smyth et al. 2001; Mashayek et al. 2017; Ijichi and Hibiya 2018; Ijichi et al. 2020). Some of those results match well with a simple relationship derived from theoretical scaling (Ijichi and Hibiya 2018), ΓROT4/3, whereas our results indicate ΓROT0.4, presenting a gentler decreasing trend. However, if we focus on the ranges for ROT < 0.2 and ROT > 1, Γ decreases with ROT with a slope close to −4/3. It is notable that for 0.2 < ROT < 1, Γ changes little with ROT with a slightly increasing tendency. The scenario in region E is similar to that in region Q (Figs. 10a,c), but the quantitative discrepancies exist, such as the fitting curve is with (Γ*, Reb*) = (0.3, 40), ΓROT0.5 in the whole ROT range, and slightly varied slope in different ranges of ROT.

What should be noted is that the turbulent patches with greater ROT correspond to higher Reb (Figs. 9a,b), and also patches with greater Reb correspond to higher ROT (Figs. 9c,d), implying Reb and ROT are not independent but positively correlated. A greater ROT indicates the turbulence is in an “elder” stage of its life-span, which suggests stronger dissipation and weaker stratification, and hence greater Reb. However, some earlier studies suggested no correlation between Reb and ROT (Mater et al. 2015). The covariation of Reb and ROT here suggests they both have an impact on the variation of Γ; therefore, the turbulent patches are classified into different ROT (Reb) groups to explore the dependence of Γ on Reb (ROT) (Figs. 9b,d). A clear monotonic dependence of Γ on Reb or ROT appears when only considering turbulent patches within fixed ROT and Reb groups (Figs. 9b,d), and both of them can be well approximated by power-law relations in the forms as Γc1Reba and Γc2ROTb, where c1 and c2 are scaling factors; the fitted values of c1, c2, a, and b are presented in detail in Tables 1 and 2.

Table 1

Coefficients for least squares fitted relations between Γ and Reb within different ranges of ROT in regions Q and E. Fittings presenting slopes distinct from 1/2 or −4/3 are marked by bold and italic font.

Table 1
Table 2

Coefficients for least squares fitted relations between Γ and ROT within different ranges of Reb in regions Q and E. Fitting presenting slope distinct from 1/2 or −4/3 is marked by bold and italic font.

Table 2

In contrast to Fig. 9a, Γ presents a remarkable increasing trend with increasing Reb for ROT < 0.2, 0.5 ≤ ROT ≤ 1 and 0.5 ≤ ROT ≤ 1 (Fig. 9b). This positive dependence of Γ on Reb is consistent with the results reported by Mashayek et al. (2017) and Ijichi et al. (2020) when Reb is relatively small but contrast to the conclusions given by Monismith et al. (2018) and Vladoiu et al. (2021). The exponent a ≈ 1/2 for the three groups with ROT < 0.2, 0.2 ≤ ROT ≤ 0.5 and 0.5 ≤ ROT ≤ 1, same as Γ ∝ Reb1/2 when Reb < Reb* reported by Mashayek et al. (2017), and c1 values are 0.16, 0.06, and 0.04, decreasing gradually with increasing ROT. The variability with ROT > 1 deviates from this trend largely due to the relatively few patches in that range.

As for the relationship between Γ and ROT, the fitted exponent b varies between −1.1 and −1.35 for the four different Reb groups, showing the −4/3 law works well in region Q. The increasing Reb alters the scaling factor c2, which increases from 0.01 to 0.13. Similar results are obtained in region E after removing some largely biased scatters (Figs. 10b,d; Tables 1 and 2). Overall, Γ is positively related to Reb as Γ ∝ Reb1/2 and negatively related to ROT as Γ ∝ ROT4/3 in both regions. These opposite monotonic relations act together to contribute to the nonmonotonic variation of Γ with Reb or ROT in the upper SCS.

The term Γ in regions Q and E share similar dependences on Reb or ROT, suggesting the variation of Γ does not change much between the energetic and quiescent regions. This is not consistent with the results in the western Mediterranean Basin reported in Vladoiu et al. (2021), where Γ showed sharply contrasted dependences on Reb in the energetic and quiescent regions: Γ decreased linearly (in logarithmic space) in the quiescent region but remained constant for Reb < 104 and decreased for Reb > 104 in the energetic region. Combining the dependences of Γ on both Reb and ROT in regions Q and E, we infer ΓReb1/2ROT4/3 in the upper SCS. This is verified by the variations of the median values of Γ binned by both Reb and ROT in regions Q and E (Fig. 11). As shown in Fig. 11, Γ increases with increasing Reb and decreasing ROT, corresponding to the positive exponent a and negative exponent b; the decreasing rate of Γ with ROT is much greater than the increasing rate of Γ with Reb, consistent with |b| being almost three times larger than |a|. The isolines of Reb1/2ROT4/3 generally match well with the Γ variations for ROT > 0.05 for both regions. The large values of Γ with ROT < 0.05 are mostly contributed by the turbulence patches in the initial stage, and whether this relation holds well for them or not should be tested using more data. Here, Γ0.020Reb1/2ROT4/3 for region E, and Γ0.016Reb1/2ROT4/3 for region Q, consistent with Γ being slightly greater in region E than that in region Q.

Fig. 11.
Fig. 11.

Variations of bin-median Γ with Reb and ROT in regions (a) E and (b) Q. The gray dashed curves are isolines of Reb1/2ROT4/3, and the gray dotted line marks ROT = 0.05.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

4. Discussion

a. Temporal variation of Γ influenced by internal wave

Between 2130 LT 10 May and 0200 LT 13 May 2010, a station in the central SCS (black circle in Fig. 1) continuously took measurements at a sampling interval of approximately 1 h, and the temporal variation of Γ over 50 h is evaluated based on these measurements. The depth-temporal variation of the median value of Γ in each 50-m-depth bin is fragmented (Fig. 12a), and the temporal variation of the median value of Γ for each cast suggests Γ at this station is mostly weaker than 0.2 (Fig. 12b). The most striking is Γ dropped sharply from 1600 LT 11 May to 1200 LT 12 May, corresponding to the period from the 18.5th to 38.5th hour. The median value of Γ dropped to 0.1 or below between 0000 LT 11 May and 0800 LT 12 May (between the 26.5th and 34.5th hour).

Fig. 12.
Fig. 12.

Temporal variations of the median value of Γ binned by (a) depth (50 m) and time (5 h), (b) median value of Γ of upper 310 m binned by cast, (c) median value of Reb with depth between 50 and 150 m binned by cast, (d) median value of ROT with depth between 150 and 310 m binned by cast, (e) log10(ε) [W kg−1; shading] with contours of potential density σ − 1000 [kg m−3; black curves] and (f) N2 (s−2) in different layers at the station in the central SCS (black circle in Fig. 1) between 2130 LT 10 May and 0200 LT 13 May 2010. The gray dots in (a) denote turbulent patches observed at this station. The black dots in (b)–(d) are the median value of Γ (Reb, ROT) binned by cast, and the red line is the result smoothed with a 3-h moving average. The two black dashed lines in (b) denote Γ = 0.2 and Γ = 0.1, respectively. The white lines in (a) and (e) and the blue lines in (b)–(f) mark the period when Γ drops sharply. Black dots in the top of (e) mark observation timings.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

Using the velocity data obtained by a ship-mounted ADCP and wavelet analysis, Liang et al. (2017) suggested there were energetic high-frequency internal waves in the thermocline in the same duration at the same location, associated with strong vertical shear of horizontal velocity and elevated TKE dissipation rate. The temporal variation of isopycnals indicated the occurrence of high-frequency internal waves, that is, the isopycnals were sharply perturbated vertically at high frequency through the water column from 50 to 310 m and were more uneven before and after this period (Fig. 12e). The elevated ε reaching 10−8 W kg−1 shallower than 150 m from the 18.5th to 38.5th hour was caused by the high-frequency internal waves and also likely by other factors (Fig. 12e), which were responsible for the decreasing Γ here as dissipation consumed a larger fraction of TKE. Jointly influenced by weakened stratification, median value of Reb increased from 9 before the 18.5th hour to 28 between the 18.5th and 38.5th hour (Figs. 12c,f). Turbulent mixing driven by high-frequency internal waves weakened the stratification from 1 × 10−4 to 7 × 10−5 s−2; however, in this process, the stratification of the upper water (above 150 m) was weakened, whereas that of the lower water (below 150 m) was strengthened (Fig. 12f). Ijichi and Hibiya (2018) noted that stratification may influence the time required for turbulence to reach the mature stage, and a stronger stratification can shorten the young stage of turbulence. As a result, the strengthened stratification at depth deeper than 150 m increased ROT by 2 times (Fig. 12d) and reduced Γ to a minimum of 0.06 (Fig. 12b). This weak Γ induced by high-frequency internal waves is also consistent with the implication of Lelong and Dunkerton (1998), that is, turbulent mixing driven by high-frequency internal waves is less efficient.

b. Estimating the Thorpe-scale-method inferred ε

The Thorpe-scale method can infer ε as ε=LO2N3=ROT2LT2N3 using the density/temperature profiles obtained from conventional hydrographic measurements, like CTD, which are relatively accessible and have been accumulated over decades, when compared with microscale measurements. Besides, the Thorpe-scale method has little constraint compared with some mixing parameterizations based on nonlinear wave–wave interaction of internal waves, so it has been widely used to explore the spatial distribution and temporal evolution of ε and/or Kρ in the open ocean (e.g., Thompson et al. 2007) and in marginal seas, including the SCS (Alford et al. 2011; Yang et al. 2016; Lu et al. 2021). Moreover, the Thorpe-scale method is employed in numerical simulations to parameterize lee wave–induced diapycnal mixing (Klymak and Legg 2010; Buijsman et al. 2012).

Although a linear relationship between LT and LO was validated in the statistical sense in various regions (Dillon 1982; Crawford 1986; Stansfield et al. 2001), indicating ROT can be treated as a constant on the order of one, which was shown to be statistically valid by comparing with direct microscale measurements (e.g., Dillon 1982; Alford et al. 2011), our results, along with other field experiments (e.g., Smyth et al. 2001; Mater et al. 2015; Ijichi and Hibiya 2018), all suggest that the relationship between LT and LO is more complicated than a linear correlation, and ROT varies too much to be simplified as a constant. Treating ROT as a constant would inevitably lead to errors of ε inferred from the Thorpe-scale method (Mater et al. 2015; Scotti 2015; Ijichi and Hibiya 2018).

Here, we evaluate the applicability of the Thorpe-scale method in the upper SCS with ROT being assigned to two different forms: 1) constant ROT (cROT), cROT = 1, representing the conventional practice of the Thorpe-scale method, and 2) parameterized ROT (pROT), pROT = 10−0.002z, where z is the depth of the overturn and 0.002 (m−1) is the observed mean decreasing rate of ROT with depth. The parameterization in form 2 suggests ROT logarithmically decreases with depth from O(1) at the surface as we observed in the upper SCS (Fig. 6f). The applicability of the Thorpe-scale method is estimated by the difference between the ε inferred from microscale observations, εOBS, and the ε inferred from the Thorpe-scale method with different ROT, εcROT and εpROT, respectively, quantified by the magnitude difference between them as Δm=|log10(εOBS)log10(εcROT)| or Δm=|log10(εOBS)log10(εpROT)|.

When applying cROT = 1 to the Thorpe-scale method, as in Figs. 13 and 13c, εOBS greater than 3.2 × 10−10 W kg−1 is relatively well parameterized by εcROT with mean Δm of 0.76 in region Q and 0.57 in region E, suggesting the biases of most patches are smaller than an order of magnitude. However, εOBS smaller than 3.2 × 10−10 W kg−1 is basically overestimated by the Thorpe-scale method with corresponding mean Δm of 1.54 in region Q and 1.34 in region E; the smaller the observed value, the greater the overestimation. This inconsistency comes from the decreasing of both ε and ROT in the upper SCS (Figs. 6b,f), the large ε in the shallow water can be fairly parameterized by ROT = 1, while the small ε at depth tends to correspond to small ROT, where ROT = 1 cannot properly scale LT to LO but overestimates LO, leading to overestimated parameterization results. These suggest the Thorpe-scale method with constant ROT tends to overestimate ε to a larger extent with depth, consistent with previous findings (Mater et al. 2015; Ijichi and Hibiya 2018). Quantitively, taking ROT as a constant leads nonnegligible bias to the Thorpe-scale-method inferred ε: the mean Δm is 0.90 in region E and 1.33 in region Q, and εcROT of 41.2% (69.7%) of turbulent patches in region Q (E) deviates from εOBS by more than an order of magnitude. This bias leads to a weak correlation between εcROT and εOBS, and the correlation coefficient is only 0.24 (0.29) in region Q (E).

Fig. 13.
Fig. 13.

Comparison between ε (W kg−1) inferred from microscale observations, εOBS, and that inferred from the Thorpe-scale method with different ROT, (a),(c) εcROT and (b),(d) εpROT in regions (top) E and (bottom) Q; εcROT is estimated with ROT = 1; εpROT with ROT = 10−0.002z. The gray dashed line marks ε = 3.2 × 10−10 W kg−1; the black line denotes εOBS = εcROT (εpROT), while the black dashed lines indicate they are biased by an order of magnitude.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

To relieve the overestimation in the deep ocean due to improper ROT, we use a logarithmically decreasing-with-depth pROT instead (Figs. 13b,d). The performance of the Thorpe-scale method for εOBS smaller than 3.2 × 10−10 W kg−1 is significantly improved with the mean Δm reduced to 0.66 in region Q and 0.64 in region E; this depth-varying ROT only slightly influences the estimate for εOBS greater than 3.2 × 10−10 W kg−1, and the corresponding mean Δm values are 0.70 and 0.56 in regions Q and E, respectively. As a result, the mean Δm is reduced to 0.67 in region Q and 0.61 in region E, and the proportions of patches with Δm ≥ 1 are significantly lowered compared with those when taking ROT = 1, being 18.9% in region E and 20.7% in region Q. And the correlation between εpROT and εOBS is enhanced as expected, with the correlation coefficient elevated from 0.24 to 0.45 in region Q and 0.29 to 0.47 in region E. This suggests the applicability of the Thorpe-scale method can be remarkably improved by taking the vertical variation of ROT into account. However, our results only observed the decreasing of ROT with depth in the upper SCS, and the ubiquity of the vertical decreasing of ROT in the deep layer and other regions needs to be validated by more microscale measurements.

5. Summary

The dissipation flux coefficient Γ of turbulent, stratified flows has long been a controversial topic and is explored by various methods including DNS, laboratory experiments, and microscale observations in different regions. Yet, no systematic evaluation of mixing efficiency is conducted in the SCS, the largest marginal sea of the western Pacific. Based on 233 casts of microscale measurements obtained from 158 stations, we looked into the spatiotemporal variation of Γ in the upper SCS. To derive Γ more accurately, we employed a modified method, namely, the difference integration method, to calculate ε and χT, which gives a better performance than the conventional iterative method, especially when the observed spectrum does not match the theoretical spectrum.

We found Γ in the upper SCS is highly variable, which scatters over three orders of magnitude from 10−2 to 101. We compared Γ and other turbulence-related variables in the energetic region (region E, west of the Luzon Strait) and quiescent region (region Q, the rest of the study area deeper than 200 m) in the upper SCS and found both regions share similar stratification patterns but are distinguished by dissipation rates of TKE and thermal variance, with those in region E significantly stronger than those in region Q. These lead to larger Reb, ROT, and Γ in region E than in region Q, indicating the turbulence in the energetic region is more efficient. Vertically, Γ presents clear increasing tendency in both regions, with a greater increasing rate in region E than in region Q. The terms ε, χT, N2, Reb and, especially, ROT all decrease with depth. The decrease of ROT with depth suggests that the increase in Γ is due to increased prevalence of high-efficiency coherent structures such as young Kelvin–Helmholtz billows.

We found 75 turbulent patches with ROT < 0.05 within the thermocline. Compared with patches in the young stage with ROT < 0.5, these patches are featured with smaller ROT, smaller ε, and larger Γ, indicating they are even younger, and we defined them as initial stage, which is confirmed by their uncomplicated temperature structures. Most of these initial-stage patches were observed in region Q, and we infer it is due to the relatively mild energy supply and experiencing a longer evolution period. Moreover, we explored the dependences of Γ on Reb and ROT and found Γ is positively related to Reb and negatively related to ROT in the upper SCS, and these opposite monotonic relations act together to contribute to the nonmonotonic variations of Γ with Reb and ROT. A simple scaling was then given as ΓReb1/2ROT4/3. The decreasing Γ due to the occurrence of high-frequency internal waves was analyzed in the central SCS, which modulates Γ by altering the local TKE dissipation rate and stratification.

The vertical decreasing of ROT calls for the assessment of the applicability of the Thorpe-scale method in the upper SCS. The Thorpe-scale method tends to overestimate ε with increasing depth when taking ROT = 1, and this bias was largely relieved using a varying ROT parameterized as ROT = 10−0.002z, which suggests ROT logarithmically decreases with depth from O(1) at the surface, as we observed in the upper SCS. This modification of ROT reduced the bias between the ε inferred from microscale measurements and that from the Thorpe-scale method and improved their correlations. Moreover, the vertical decreasing of ROT is not an isolated incident in the upper SCS, since Ijichi and Hibiya (2018) found mixing events with Γ ∼ O(1) and ROTO(0.1) are more frequently observed in the deep ocean, whereas those with Γ ∼ O(0.1) and ROTO(1) tend to occur in the upper ocean in various regions of the North Pacific and Southern Oceans, suggesting the vertical decreasing of ROT is widespread in the ocean. It is notable that the scaling relations of Γ with Reb and ROT, the parameterized ROT, and the spatiotemporal variation of Γ presented in this study are all in the upper 550 m of the SCS; as a result, more microstructure samplings and associated research in the deep SCS should be expected.

Acknowledgments.

This work was supported by the National Natural Science Foundation of China (Grants 42076012, 42006012, 41676008, and 92158201), the Natural Science Outstanding Youth Fund of Shandong Province (ZR2019JQ13), the College Innovation Team Project of Guangdong Province (2019KCXTF021), and the National Key Research and Development Program of China (2017YFA0604102).

Data availability statement.

The bathymetry data used in this study are available at https://ngdc.noaa.gov/mgg/global/global.html. The microstructure data are available upon request to corresponding authors.

APPENDIX

Comparison of Methods and Their Influences on Results

A new method, the difference integration method, is used to derive ε and χT in this study. We examine the differences between this method and the iterative method and their influences on the results presented in the main text. Fig. A1 presents two cases with “normal” and “anomalous” spectra when we employ two different methods. Here, normal means the measured spectra can be well fitted by the Nasmyth or Bachelor spectrum with adjustment, while the anomalous spectra keep on departing significantly from the Nasmyth or Bachelor spectrum after numerous adjustments. The subscript “D” represents the difference integration method and “I” for the iterative method. For the case shown in Fig. A1a, the normal shear spectrum can be well fitted by both methods; the resulting ε are quite similar, with εI of 8.0 × 10−9 W kg−1 and εD of 1.3 × 10−8 W kg−1. However, for the case in Fig. A1b, the measured shear spectrum is anomalous, and the spectrum in low wavenumbers uplifts significantly, making it tricky to obtain a good fitting for the iterative method. Especially, the result of the iterative method is seriously influenced by the uplift, resulting in a large εI that is about seven times as large as εD. Moreover, the integrated difference between the measured shear spectrum and the Nasmyth spectrum of the new method, DφD, is only half of that of the iterative method DφI. Considering all the spectra, we can see, for the normal shear spectrum, εI is slightly greater than εD, with εI/εD no greater than 4, while for the anomalous spectrum, εI is about an order of magnitude greater than εD (i.e., εI/εD ≥ 6). The scenario for temperature gradient spectra is similar.

Fig. A1.
Fig. A1.

Cases indicating the difference of ε calculated by two different methods, (a) normal shear spectra, which can be well fitted by Nasmyth spectra, and (b) anomalous shear spectra. Subscript “D” represents the difference integration method and “I” represents the iterative method. The results and integrated difference between spectra are noted in (a) and (b). (c) The proportions of spectra with εI/εD (χTI/χTD) varying from 2 to 10.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

Since ε estimated from the anomalous spectrum is less credible, it is necessary to examine the frequency of occurrence of the anomalous spectra; εI/εD ≤ 4 (χTI/χTD ≤ 4) and εI/εD ≥ 6 (χTI/χTD ≥ 6) are used as indicators to distinguish normal and anomalous spectra. As shown in Fig. A1c, we measured the proportions of spectra with εI/εD varying from 2 to 10. More than 90% of the spectra have εI/εD ≤ 4, suggesting most shear spectra are normal and can be well fitted by both methods. Only fewer than 5% of the spectra have εI/εD ≥ 6, indicating anomalous spectra are only a small fraction. The situation for temperature gradient spectra and χT is similar to that for shear spectra and ε (magenta curve in Fig. A1c).

Next, the differences of ε and χT resulted from the two methods are estimated. Fig. A2a compares εD and εI; εI is greater than εD by about 1.8 times on average. A greater εI leads to a greater Batchelor wavenumber kBI, so χTI is integrated as χTI=6κ0kBIψ(k)dk in a broader wavenumber range and is also greater than χTD by about 2.2 times (Fig. A2b). So, both ε and χT calculated by the iterative method are about twice as large as those calculated by the difference integration method. The differences of ε and χT resulted from the two methods require us to assess which method is closer to the truth. We use the integrated differences between the shear spectrum and the Nasmyth spectra corresponding to εI and εD, DφI and DφD, as indications of which method is closer to the truth. If DφI > DφD, it suggests the Nasmyth spectrum fitted by the iterative method is more deviated from the shear spectrum than that of the difference integration method, so the latter is closer to the truth; for DφI < DφD, the result of the iterative method is closer to the truth. The comparison between DφI and DφD of all shear spectra suggests 94.1% of the spectra have DφI > DφD (Fig. A2c), with DφI/DφD being 2.4 on average, suggesting ε calculated by the difference integration method is closer to the truth. And for the temperature gradient spectra and χT, DψI and DψD are compared (Fig. A2d). We find that 94.3% of the spectra have DψI > DψD, and DψI/DψD is 2.3 on average, which also suggests the result from the difference integration method is closer to the truth.

Fig. A2.
Fig. A2.

Comparisons between (a) εD and εI, (b) χTD and χTI, (c) DφI and DφD, and (d) DψI and DψD.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

Moreover, we examined the influences of different methods on the conclusions drawn in this paper. For Γ, since εI and χTI are both about twice as large as εD and χTD, ΓI is only slightly greater than ΓD (Fig. A3a), the median values are 0.21 and 0.19, respectively. The variable ΓI also presents a clear increasing trend with depth, though its increasing rate is smaller than that of ΓD; Reb and ROT derived from the iterative method are both increased compared with those from the difference integration method, as the consequence of larger εI than εD. And their median values are amplified for about 1.7 and 1.4, respectively. Their vertical variations remain unchanged after using the iterative method (Figs. A3b,c). With these minor differences, the relationship between ΓI, RebI, and ROTI can still be scaled as ΓReb1/2ROT4/3 (Fig. A3d), which is consistent with the scaling based on the difference integration method. These suggest the results and conclusions based on these two methods are consistent, since the new method only performs better when the spectra are anomalous, which is rare in this study.

Fig. A3.
Fig. A3.

Comparison of (a) Γ, (b) Reb, and (c) ROT derived from different methods, and (d) the variation of Γ with Reb and ROT based on the iterative method. The fitted slopes of Γ ∝ Reb and Γ ∝ ROT are noted in the right panels in (d), which are consistent with those based on the difference integration method.

Citation: Journal of Physical Oceanography 53, 2; 10.1175/JPO-D-22-0127.1

REFERENCES

  • Alford, M. H., and Coauthors, 2011: Energy flux and dissipation in Luzon Strait: Two tales of two ridges. J. Phys. Oceanogr., 41, 22112222, https://doi.org/10.1175/JPO-D-11-073.1.

    • Search Google Scholar
    • Export Citation
  • Barry, M. E., G. N. Ivey, K. B. Winters, and J. Imberger, 2001: Measurements of diapycnal diffusivities in stratified fluids. J. Fluid Mech., 442, 267291, https://doi.org/10.1017/S0022112001005080.

    • Search Google Scholar
    • Export Citation
  • Bouruet-Aubertot, P., C. Koudella, C. Staquet, and K. B. Winters, 2001: Particle dispersion and mixing induced by breaking internal gravity waves. Dyn. Atmos. Oceans, 33, 95134, https://doi.org/10.1016/S0377-0265(00)00056-7.

    • Search Google Scholar
    • Export Citation
  • Buijsman, M. C., S. Legg, and J. Klymak, 2012: Double-ridge internal tide interference and its effect on dissipation in Luzon Strait. J. Phys. Oceanogr., 42, 13371356, https://doi.org/10.1175/JPO-D-11-0210.1.

    • Search Google Scholar
    • Export Citation
  • Chor, T., J. O. Wenegrat, and J. Taylor, 2022: Insights into the mixing efficiency of submesoscale centrifugal-symmetric instabilities. J. Phys. Oceanogr., 52, 22732287, https://doi.org/10.1175/JPO-D-21-0259.1.

    • Search Google Scholar
    • Export Citation
  • Cimoli, L., C. P. Caulfield, H. L. Johnson, D. P. Marshall, A. Mashayek, A. C. Naveira Garabato, and C. Vic, 2019: Sensitivity of deep ocean mixing to local internal tide breaking and mixing efficiency. Geophys. Res. Lett., 46, 14 62214 633, https://doi.org/10.1029/2019GL085056.

    • Search Google Scholar
    • Export Citation
  • Crawford, W. R., 1986: A comparison of length scales and decay times of turbulence in stably stratified flows. J. Phys. Oceanogr., 16, 18471854, https://doi.org/10.1175/1520-0485(1986)016<1847:ACOLSA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Dillon, T. M., 1982: Vertical overturns: A comparison of Thorpe and Ozmidov length scales. J. Geophys. Res., 87, 96019613, https://doi.org/10.1029/JC087iC12p09601.

    • Search Google Scholar
    • Export Citation
  • Fernando, H. J. S., 1991: Turbulent mixing in stratified fluids. Annu. Rev. Fluid Mech., 23, 455493, https://doi.org/10.1146/annurev.fl.23.010191.002323.

    • Search Google Scholar
    • Export Citation
  • Gregg, M. C., and J. K. Horne, 2009: Turbulence, acoustic backscatter, and pelagic nekton in Monterey Bay. J. Phys. Oceanogr., 39, 10971114, https://doi.org/10.1175/2008JPO4033.1.

    • Search Google Scholar
    • Export Citation
  • Gregg, M. C., E. A. D’Asaro, J. J. Riley, and E. Kunze, 2018: Mixing efficiency in the ocean. Annu. Rev. Mar. Sci., 10, 443473, https://doi.org/10.1146/annurev-marine-121916-063643.

    • Search Google Scholar
    • Export Citation
  • Holleman, R. C., W. R. Geyer, and D. K. Ralston, 2016: Stratified turbulence and mixing efficiency in a salt wedge estuary. J. Phys. Oceanogr., 46, 17691783, https://doi.org/10.1175/JPO-D-15-0193.1.

    • Search Google Scholar
    • Export Citation
  • Ijichi, T., and T. Hibiya, 2018: Observed variations in turbulent mixing efficiency in the deep ocean. J. Phys. Oceanogr., 48, 18151830, https://doi.org/10.1175/JPO-D-17-0275.1.

    • Search Google Scholar
    • Export Citation
  • Ijichi, T., L. St. Laurent, K. L. Polzin, and J. M. Toole, 2020: How variable is mixing efficiency in the abyss? Geophys. Res. Lett., 47, e2019GL086813, https://doi.org/10.1029/2019GL086813.

    • Search Google Scholar
    • Export Citation
  • Ivey, G. N., and J. Imberger, 1991: On the nature of turbulence in a stratified fluid. Part I: The energetics of mixing. J. Phys. Oceanogr., 21, 650658, https://doi.org/10.1175/1520-0485(1991)021<0650:OTNOTI>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Jackson, P. R., and C. R. Rehmann, 2003: Laboratory measurements of differential diffusion in a diffusively stable, turbulent flow. J. Phys. Oceanogr., 33, 15921603, https://doi.org/10.1175/2405.1.

    • Search Google Scholar
    • Export Citation
  • Jackson, P. R., and C. R. Rehmann, 2014: Experiments on differential scalar mixing in turbulence in a sheared, stratified flow. J. Phys. Oceanogr., 44, 26612680, https://doi.org/10.1175/JPO-D-14-0027.1.

    • Search Google Scholar
    • Export Citation
  • Klymak, J. M., and S. M. Legg, 2010: A simple mixing scheme for models that resolve breaking internal waves. Ocean Modell., 33, 224234, https://doi.org/10.1016/j.ocemod.2010.02.005.

    • Search Google Scholar
    • Export Citation
  • Klymak, J. M., M. H. Alford, R. Pinkel, R.-C. Lien, Y. J. Yang, and T.-Y. Tang, 2011: The breaking and scattering of the internal tide on a continental slope. J. Phys. Oceanogr., 41, 926945, https://doi.org/10.1175/2010JPO4500.1.

    • Search Google Scholar
    • Export Citation
  • Lelong, M.-P., and T. J. Dunkerton, 1998: Inertia–gravity wave breaking in three dimensions. Part I: Convectively stable waves. J. Atmos. Sci., 55, 24732488, https://doi.org/10.1175/1520-0469(1998)055<2473:IGWBIT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Liang, C.-R., G.-Y. Chen, and X.-D. Shang, 2017: Observations of the turbulent kinetic energy dissipation rate in the upper central South China Sea. Ocean Dyn., 67, 597609, https://doi.org/10.1007/s10236-017-1051-6.

    • Search Google Scholar
    • Export Citation
  • Lu, Y. Z., S. Zhou, X. Cen, S. Guo, and X. Shang, 2014: Salt finger and turbulent mixing in the upper layer of the central southern South China Sea. Oceanol. Limnol. Sin., 45, 11581167.

    • Search Google Scholar
    • Export Citation
  • Lu, Y. Z., X.-R. Cen, S.-X. Guo, L. Qu, P.-Q. Huang, X.-D. Shang, and S.-Q. Zhou, 2021: Spatial variability of diapycnal mixing in the South China Sea inferred from density overturn analysis. J. Phys. Oceanogr., 51, 34173434, https://doi.org/10.1175/JPO-D-20-0241.1.

    • Search Google Scholar
    • Export Citation
  • Mashayek, A., H. Salehipour, D. Bouffard, C. P. Caulfield, R. Ferrari, M. Nikurashin, W. R. Peltier, and W. D. Smyth, 2017: Efficiency of turbulent mixing in the abyssal ocean circulation. Geophys. Res. Lett., 44, 62966306, https://doi.org/10.1002/2016GL072452.

    • Search Google Scholar
    • Export Citation
  • Masunaga, E., S. K. Venayagamoorthy, K. Wada, and H. Yamazaki, 2022: Variability of the diapycnal mixing coefficient in coastal oceans investigated with direct microstructure measurements. J. Mar. Syst., 230, 103722, https://doi.org/10.1016/j.jmarsys.2022.103722.

    • Search Google Scholar
    • Export Citation
  • Mater, B. D., S. K. Venayagamoorthy, L. St. Laurent, and J. N. Moum, 2015: Biases in Thorpe-scale estimates of turbulence dissipation. Part I: Assessments from large-scale overturns in oceanographic data. J. Phys. Oceanogr., 45, 24972521, https://doi.org/10.1175/JPO-D-14-0128.1.

    • Search Google Scholar
    • Export Citation
  • Monismith, S. G., J. R. Koseff, and B. L. White, 2018: Mixing efficiency in the presence of stratification: When is it constant? Geophys. Res. Lett., 45, 56275634, https://doi.org/10.1029/2018GL077229.

    • Search Google Scholar
    • Export Citation
  • Moum, J. N., 1996a: Efficiency of mixing in the main thermocline. J. Geophys. Res., 101, 12 05712 069, https://doi.org/10.1029/96JC00508.

    • Search Google Scholar
    • Export Citation
  • Moum, J. N., 1996b: Energy-containing scales of turbulence in the ocean thermocline. J. Geophys. Res., 101, 14 09514 109, https://doi.org/10.1029/96JC00507.

    • Search Google Scholar
    • Export Citation
  • Nasmyth, P. W., 1970: Oceanic turbulence. Ph.D. thesis, Institute of Oceanography, University of British Columbia, 69 pp.

  • Niwa, Y., and T. Hibiya, 2004: Three-dimensional numerical simulation of M2 internal tides in the East China Sea. J. Geophys. Res., 109, C04027, https://doi.org/10.1029/2003JC001923.

    • Search Google Scholar
    • Export Citation
  • Oakey, N. S., 1982: Determination of the rate of dissipation of turbulent energy from simultaneous temperature and velocity shear microstructure measurements. J. Phys. Oceanogr., 12, 256271, https://doi.org/10.1175/1520-0485(1982)012<0256:DOTROD>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Osborn, T. R., 1980: Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr., 10, 8389, https://doi.org/10.1175/1520-0485(1980)010<0083:EOTLRO>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Osborn, T. R., and C. S. Cox, 1972: Oceanic fine structure. Geophys. Fluid Dyn., 3, 321345, https://doi.org/10.1080/03091927208236085.

    • Search Google Scholar
    • Export Citation
  • Ozmidov, R. V., 1965: On the turbulent exchange in a stably stratified ocean. Izv. Acad. Sci. USSR, Atmos. Oceanic Phys., 1, 861871.

  • Qu, T., J. B. Girton, and J. A. Whitehead, 2006: Deepwater overflow through Luzon Strait. J. Geophys. Res., 111, C01002, https://doi.org/10.1029/2005JC003139.

    • Search Google Scholar
    • Export Citation
  • Ruddick, B., 1983: A practical indicator of the stability of the water column to double-diffusive activity. Deep-Sea Res., 30A, 11051107, https://doi.org/10.1016/0198-0149(83)90063-8.

    • Search Google Scholar
    • Export Citation
  • Ruddick, B., D. Walsh, and N. Oakey, 1997: Variations in apparent mixing efficiency in the North Atlantic central water. J. Phys. Oceanogr., 27, 25892605, https://doi.org/10.1175/1520-0485(1997)027<2589:VIAMEI>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Salehipour, H., C. P. Caulfield, and W. R. Peltier, 2016: Turbulent mixing due to the Holmboe wave instability at high Reynolds number. J. Fluid Mech., 803, 591621, https://doi.org/10.1017/jfm.2016.488.

    • Search Google Scholar
    • Export Citation
  • Scotti, A., 2015: Biases in Thorpe-scale estimates of turbulence dissipation. Part II: Energetics arguments and turbulence simulations. J. Phys. Oceanogr., 45, 25222543, https://doi.org/10.1175/JPO-D-14-0092.1.

    • Search Google Scholar
    • Export Citation
  • Seim, H. E., and M. C. Gregg, 1994: Detailed observations of a naturally occurring shear instability. J. Geophys. Res., 99, 10 04910 073, https://doi.org/10.1029/94JC00168.

    • Search Google Scholar
    • Export Citation
  • Shang, X.-D., C.-R. Liang, and G.-Y. Chen, 2017: Spatial distribution of turbulent mixing in the upper ocean of the South China Sea. Ocean Sci., 13, 503519, https://doi.org/10.5194/os-13-503-2017.

    • Search Google Scholar
    • Export Citation
  • Shih, L. H., J. R. Koseff, G. N. Ivey, and J. H. Ferziger, 2005: Parameterization of turbulent fluxes and scales using homogeneous sheared stably stratified turbulence simulations. J. Fluid Mech., 525, 193214, https://doi.org/10.1017/S0022112004002587.

    • Search Google Scholar
    • Export Citation
  • Smyth, W. D., 2020: Marginal instability and the efficiency of ocean mixing. J. Phys. Oceanogr., 50, 21412150, https://doi.org/10.1175/JPO-D-20-0083.1.

    • Search Google Scholar
    • Export Citation
  • 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, 19691992, https://doi.org/10.1175/1520-0485(2001)031<1969:TEOMIT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Smyth, W. D., J. R. Carpenter, and G. A. Lawrence, 2007: Mixing in symmetric Holmboe waves. J. Phys. Oceanogr., 37, 15661583, https://doi.org/10.1175/JPO3037.1.

    • Search Google Scholar
    • Export Citation
  • St. Laurent, L., 2008: Turbulent dissipation on the margins of the South China Sea. Geophys. Res. Lett., 35, L23615, https://doi.org/10.1029/2008GL035520.

    • Search Google Scholar
    • Export Citation
  • St. Laurent, L., and R. W. Schmitt, 1999: The contribution of salt fingers to vertical mixing in the North Atlantic tracer release experiment. J. Phys. Oceanogr., 29, 14041424, https://doi.org/10.1175/1520-0485(1999)029<1404:TCOSFT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • St. Laurent, L., H. L. Simmons, and S. R. Jayne, 2002: Estimating tidally driven mixing in the deep ocean. Geophys. Res. Lett., 29, 2106, https://doi.org/10.1029/2002GL015633.

    • Search Google Scholar
    • Export Citation
  • Stansfield, K., C. Garrett, and R. Dewey, 2001: Probability distribution of the Thorpe displacement within overturns in Juan de Fuca Strait. J. Phys. Oceanogr., 31, 34213434, https://doi.org/10.1175/1520-0485(2001)031<3421:TPDOTT>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Sun, H., and Q. Wang, 2016: Microstructure observations in the upper layer of the South China Sea. J. Oceanogr., 72, 777786, https://doi.org/10.1007/s10872-016-0371-3.

    • Search Google Scholar
    • Export Citation
  • Thompson, A. F., S. T. Gille, J. A. MacKinnon, and J. Sprintall, 2007: Spatial and temporal patterns of small-scale mixing in Drake Passage. J. Phys. Oceanogr., 37, 572592, https://doi.org/10.1175/JPO3021.1.

    • Search Google Scholar
    • Export Citation
  • Thorpe, S. A., 1977: Turbulence and mixing in a Scottish loch. Philos. Trans. Roy. Soc., A286, 125181, https://doi.org/10.1098/rsta.1977.0112.

    • Search Google Scholar
    • Export Citation
  • Tian, J., Q. Yang, and W. Zhao, 2009: Enhanced diapycnal mixing in the South China Sea. J. Phys. Oceanogr., 39, 31913203, https://doi.org/10.1175/2009JPO3899.1.

    • Search Google Scholar
    • Export Citation
  • Venayagamoorthy, S. K., and J. R. Koseff, 2016: On the flux Richardson number in stably stratified turbulence. J. Fluid Mech., 798, R1, https://doi.org/10.1017/jfm.2016.340.

    • Search Google Scholar
    • Export Citation
  • Vladoiu, A., P. Bouruet-Aubertot, Y. Cuypers, B. Ferron, K. Schroeder, M. Borghini, and S. Leizour, 2021: Contrasted mixing efficiency in energetic versus quiescent regions: Insights from microstructure measurements in the western Mediterranean Sea. Prog. Oceanogr., 195, 102594, https://doi.org/10.1016/j.pocean.2021.102594.

    • Search Google Scholar
    • Export Citation
  • Wijesekera, H. W., and T. M. Dillon, 1997: Shannon entropy as an indicator of age for turbulent overturns in the oceanic thermocline. J. Geophys. Res., 102, 32793291, https://doi.org/10.1029/96JC03605.

    • Search Google Scholar
    • Export Citation
  • Wolk, F., H. Yamazaki, L. Seuront, and R. G. Lueck, 2002: A new free-fall profiler for measuring biophysical microstructure. J. Atmos. Oceanic Technol., 19, 780793, https://doi.org/10.1175/1520-0426(2002)019<0780:ANFFPF>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Wunsch, C., and R. Ferrari, 2004: Vertical mixing, energy, and the general circulation of the oceans. Annu. Rev. Fluid Mech., 36, 281314, https://doi.org/10.1146/annurev.fluid.36.050802.122121.

    • Search Google Scholar
    • Export Citation
  • Yang, Q., W. Zhao, X. Liang, and J. Tian, 2016: Three-dimensional distribution of turbulent mixing in the South China Sea. J. Phys. Oceanogr., 46, 769788, https://doi.org/10.1175/JPO-D-14-0220.1.

    • Search Google Scholar
    • Export Citation
  • You, Y., 2002: A global ocean climatological atlas of the Turner angle: Implications for double-diffusion and water-mass structure. Deep-Sea Res. I, 49, 20752093, https://doi.org/10.1016/S0967-0637(02)00099-7.

    • Search Google Scholar
    • Export Citation
  • Zhao, Z., 2014: Internal tide radiation from the Luzon Strait. J. Geophys. Res. Oceans, 119, 54345448, https://doi.org/10.1002/2014JC010014.

    • Search Google Scholar
    • Export Citation
  • Zhou, C., W. Zhao, J. Tian, Q. Yang, and T. Qu, 2014: Variability of the deep-water overflow in the Luzon Strait. J. Phys. Oceanogr., 44, 29722986, https://doi.org/10.1175/JPO-D-14-0113.1.

    • Search Google Scholar
    • Export Citation
Save
  • Alford, M. H., and Coauthors, 2011: Energy flux and dissipation in Luzon Strait: Two tales of two ridges. J. Phys. Oceanogr., 41, 22112222, https://doi.org/10.1175/JPO-D-11-073.1.

    • Search Google Scholar
    • Export Citation
  • Barry, M. E., G. N. Ivey, K. B. Winters, and J. Imberger, 2001: Measurements of diapycnal diffusivities in stratified fluids. J. Fluid Mech., 442, 267291, https://doi.org/10.1017/S0022112001005080.

    • Search Google Scholar
    • Export Citation
  • Bouruet-Aubertot, P., C. Koudella, C. Staquet, and K. B. Winters, 2001: Particle dispersion and mixing induced by breaking internal gravity waves. Dyn. Atmos. Oceans, 33, 95134, https://doi.org/10.1016/S0377-0265(00)00056-7.

    • Search Google Scholar
    • Export Citation
  • Buijsman, M. C., S. Legg, and J. Klymak, 2012: Double-ridge internal tide interference and its effect on dissipation in Luzon Strait. J. Phys. Oceanogr., 42, 13371356, https://doi.org/10.1175/JPO-D-11-0210.1.

    • Search Google Scholar
    • Export Citation
  • Chor, T., J. O. Wenegrat, and J. Taylor, 2022: Insights into the mixing efficiency of submesoscale centrifugal-symmetric instabilities. J. Phys. Oceanogr., 52, 22732287, https://doi.org/10.1175/JPO-D-21-0259.1.

    • Search Google Scholar
    • Export Citation
  • Cimoli, L., C. P. Caulfield, H. L. Johnson, D. P. Marshall, A. Mashayek, A. C. Naveira Garabato, and C. Vic, 2019: Sensitivity of deep ocean mixing to local internal tide breaking and mixing efficiency. Geophys. Res. Lett., 46, 14 62214 633, https://doi.org/10.1029/2019GL085056.

    • Search Google Scholar
    • Export Citation
  • Crawford, W. R., 1986: A comparison of length scales and decay times of turbulence in stably stratified flows. J. Phys. Oceanogr., 16, 18471854, https://doi.org/10.1175/1520-0485(1986)016<1847:ACOLSA>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Dillon, T. M., 1982: Vertical overturns: A comparison of Thorpe and Ozmidov length scales. J. Geophys. Res., 87, 96019613, https://doi.org/10.1029/JC087iC12p09601.

    • Search Google Scholar
    • Export Citation
  • Fernando, H. J. S., 1991: Turbulent mixing in stratified fluids. Annu. Rev. Fluid Mech., 23, 455493, https://doi.org/10.1146/annurev.fl.23.010191.002323.

    • Search Google Scholar
    • Export Citation
  • Gregg, M. C., and J. K. Horne, 2009: Turbulence, acoustic backscatter, and pelagic nekton in Monterey Bay. J. Phys. Oceanogr., 39, 10971114, https://doi.org/10.1175/2008JPO4033.1.

    • Search Google Scholar
    • Export Citation
  • Gregg, M. C., E. A. D’Asaro, J. J. Riley, and E. Kunze, 2018: Mixing efficiency in the ocean. Annu. Rev. Mar. Sci., 10, 443473, https://doi.org/10.1146/annurev-marine-121916-063643.

    • Search Google Scholar
    • Export Citation
  • Holleman, R. C., W. R. Geyer, and D. K. Ralston, 2016: Stratified turbulence and mixing efficiency in a salt wedge estuary. J. Phys. Oceanogr., 46, 17691783, https://doi.org/10.1175/JPO-D-15-0193.1.

    • Search Google Scholar
    • Export Citation
  • Ijichi, T., and T. Hibiya, 2018: Observed variations in turbulent mixing efficiency in the deep ocean. J. Phys. Oceanogr., 48, 18151830, https://doi.org/10.1175/JPO-D-17-0275.1.

    • Search Google Scholar
    • Export Citation
  • Ijichi, T., L. St. Laurent, K. L. Polzin, and J. M. Toole, 2020: How variable is mixing efficiency in the abyss? Geophys. Res. Lett., 47, e2019GL086813, https://doi.org/10.1029/2019GL086813.

    • Search Google Scholar
    • Export Citation
  • Ivey, G. N., and J. Imberger, 1991: On the nature of turbulence in a stratified fluid. Part I: The energetics of mixing. J. Phys. Oceanogr., 21, 650658, https://doi.org/10.1175/1520-0485(1991)021<0650:OTNOTI>2.0.CO;2.

    • Search Google Scholar
    • Export Citation
  • Jackson, P. R., and C. R. Rehmann, 2003: Laboratory measurements of differential diffusion in a diffusively stable, turbulent flow. J. Phys. Oceanogr., 33, 15921603, https://doi.org/10.1175/2405.1.

    • Search Google Scholar
    • Export Citation
  • Jackson, P. R., and C. R. Rehmann, 2014: Experiments on differential scalar mixing in turbulence in a sheared, stratified flow. J. Phys. Oceanogr., 44, 26612680, https://doi.org/10.1175/JPO-D-14-0027.1.

    • Search Google Scholar
    • Export Citation
  • Klymak, J. M., and S. M. Legg, 2010: A simple mixing scheme for models that resolve breaking internal waves. Ocean Modell., 33, 224234, https://doi.org/10.1016/j.ocemod.2010.02.005.

    • Search Google Scholar
    • Export Citation
  • Klymak, J. M., M. H. Alford, R. Pinkel, R.-C. Lien, Y. J. Yang, and T.-Y. Tang, 2011: The breaking and scattering of the internal tide on a continental slope. J. Phys. Oceanogr., 41, 926945, https://doi.org/10.1175/2010JPO4500.1.

    • Search Google Scholar
    • Export Citation
  • Lelong, M.-P., and T. J. Dunkerton, 1998: Inertia–gravity wave breaking in three dimensions. Part I: Convectively stable waves. J. Atmos. Sci., 55, 24732488, https://doi.org/10.1175/1520-0469(1998)055