1. Introduction
Despite the two order of magnitude discrepancy between the molecular diffusivities of heat (DT ≃ 10–7 m2 s–1) and salt (DS ≃ 10–9 m2 s–1), it is commonly assumed that turbulent transports of heat and salt are equal. It is presumed that turbulent fluxes are dominated by the motions of the largest-scale eddies and that molecular processes (which produce the irreversible mixing) occur at a rate consistent with the large-scale turbulence production. In calculating salt fluxes, this has been a necessary assumption because it has not been possible to directly measure the turbulent flux of salt or the dissipation rate of salinity variance.
It is well known that double-diffusive processes transport heat and salt at different rates (Schmitt 1979). The unique dynamics associated with these structures are a direct consequence of the large value of the ratio DT/DS ≃ 100, but only occur when turbulence is weak and for a limited range of dT/dS. Having potentially broader implications are the poorly understood processes in turbulent mixing events that may lead to a differential transport of heat and salt (Gargett 2002). Both laboratory (Turner 1968; Altman and Gargett 1990) and numerical experiments (Merryfield et al. 1998; Holloway et al. 2001) indicate that heat is transported more effectively than salt in weak, stratified turbulence. This discrepancy in mixing efficiencies has been attributed to the incomplete mixing of salt. Since all previous research represents mixing in weak, low Reynolds number turbulence, it remains an open question as to whether the turbulent transports of heat and salt are significantly different in high Reynolds number ocean turbulence (Gargett 2002). Only scaling arguments have been available to estimate the role salinity plays in the eddy diffusivity for mass (or buoyancy) and its importance in the generation of entropy (Gregg 1984).
While estimates of KT (the eddy diffusivity for heat) using the methods of Osborn and Cox (1972) have been made over the past 30 years, KS (the eddy diffusivity for salt) has eluded measurement. In particular, it has not been possible to measure the spectrum of salinity gradient Ψ
The basic assumption that all scalars are transported equally by turbulence, regardless of their molecular diffusivities, stems from the reasoning that turbulent transports generally occur at scales much larger than the scales of molecular processes. This separation of scales is often a fundamental premise of high Reynolds number turbulence theory and analysis. Assuming the typical cascade of energy, turbulent fluctuations are produced at the scales of the largest eddies and break down into smaller and smaller eddies where they are eventually dissipated by molecular diffusion.
Of fundamental importance to the turbulent transport of salinity is the determination of the spectral shape of salinity gradient fluctuations and the dissipation rate of salinity variance. Quantification of Ψ
This paper is organized as follows. In section 2, the general form of the spectrum of salinity gradient is developed, and the experimental and analytical procedures used to obtain it are described. Observations of gradient spectra, scalar dissipation rates, and covariance fluxes are presented in section 3. In the discussion (section 4), normalized spectra are presented, and the turbulent fluxes of salinity and temperature are compared in terms of dissipation flux coefficients Γo and Γd and diffusivity ratios do and dx. We discuss how our flux estimates and the observed spectral shapes may be consistent with differential diffusion. We conclude (section 5) that the spectrum of salinity gradient has a shape similar to that of temperature gradient (except extending to higher wavenumbers) and approximately follows Kraichnan's universal form. The ratio of the eddy diffusivity of salt to that of heat is found to have an average slightly less than one. As discussed in the appendix, errors arise primarily from corrections for thermistor response and the shape of the scalar gradient spectrum at high wavenumbers; bias is also introduced due to anisotropy of weakly turbulent, buoyancy-influenced patches. As a result, the ratio dx = KS/KT is estimated to be 0.6 < dx < 1.1. While suggestive that dx < 1, these results do not rule out the possibility that KS = KT. Consequences of this study are presented in section 6.
2. Methodology
a. Theory
1) Scalar spectra
We are concerned primarily with the dissipation range of turbulence, which contains most of the variance of gradient spectra (Nash and Moum 1999). Gradients of scalars are intensified and shifted to higher wavenumbers as a result of strain by the turbulent velocity field. This cascade of energy in the inertial subrange depends only on the dissipation rate of TKE, ϵ, leading to the classic k–5/3 dependence of the velocity spectrum. Kolmogorov (1941) reasoned that viscosity only becomes important at wavenumbers of O(kη), where kη = (ϵ/ν3)1/4. At higher wavenumbers, velocity fluctuations are heavily damped by molecular viscosity.
Since the values of the molecular diffusivities of heat DT and salt DS are much smaller than that for momentum, fluctuations of T and S extend to much higher wavenumbers than fluctuations of velocity. Batchelor (1959) used this fact when he assumed that the evolution of scalars with Pr ≡ ν/Dθ ≫ 1 are governed solely by the mean least principal strain rate γ ∼ −(ϵ/ν)1/2, the strain associated with convergent motions in the turbulent velocity field.

2) Application to salinity
The following methodology differs slightly from that of Nash and Moum (1999) and Washburn et al. (1996), in which Ψ
b. Experimental details
To calculate Ψ
The μCT probe was installed on Chameleon, our loosely tethered microstructure profiler, in addition to the regular suite of microstructure sensors: a pitot tube used to measure the fluctuating vertical velocity and estimate 〈w′T′〉, 〈w′S′〉, and 〈w′2〉 (Moum 1990); airfoil shear probes used to estimate ϵ, the dissipation rate of TKE (Moum et al. 1995, e.g.); a second fast-response FP07 thermistor; and a stable Neil-Brown conductivity cell, used to calibrate the μC sensor in situ. Casts using the ship's SeaBird CTD were periodically made for comparison with Chameleon's temperature and conductivity measurements.
Because the μCT probe is designed for laboratory use, it is susceptible to fouling and damage, and it is difficult to obtain a stable laboratory calibration. As a result, we calibrate the sensor in situ: polynomial calibration coefficients are determined by fitting a low-pass filtered μC signal to the conductivity measured by the Neil-Brown cell. The coefficients obtained are then applied to the unfiltered μC signal and its derivative. Care is taken not to include μC data that contain nonphysical spikes or steplike features, which are not present in the conductivity time series from the Neil-Brown cell and likely represent an impact of the sensor with biology. After patches are selected, the μC and Neil-Brown signals are again compared; records which differ significantly are discarded.
The μC sensor was sampled at 409.6 Hz and its derivative at either 819.2 or 409.6 Hz, depending on the experiment. Thermistor temperature and its derivative were sampled at 102.4 and 204.8 Hz, respectively. Four-pole analog Butterworth filters were used for antialiasing before digitizing at 16 bits; filter cutoff frequencies of 32, 64, 132, and 245 Hz were used for signals sampled at 102.4, 204.8, 409.6, and 819.2 Hz, respectively. The transfer functions of the filters and analog differentiators were determined in the laboratory and spectral corrections to restore lost variance were applied to the data during processing.
To measure 80% of the variance of Ψ
Several hundred vertical profiles (from the surface to the bottom at depth 50–200 m) were acquired on two separate occasions on Oregon's continental shelf on the southern flanks of Heceta Bank on 23 August 1998 and over Stonewall Bank on 15 April 1999. Measurements of turbulent vertical velocity w′ were obtained only during the Heceta Bank experiment. During the Stonewall Bank experiment, a third temperature sensor (an ultrafast-response thermocouple: Nash et al. 1999) was installed on Chameleon and used as a benchmark to determine the thermistor frequency-response transfer function in situ.
The dominant currents at Heceta Bank follow local isobaths. Near the crest of the bank, the flow is mostly to the southeast and mixing is dominated by bottom-boundary processes. Offshore of the bank, the southeast flowing surface currents are opposed by a northwestward flowing undercurrent, which combine to produce an intensified shear region near depth 70 m. The stratification near the surface is mostly due to temperature; at depth, where temperature inversions and salinity intrusions are common, salinity plays a more dominant role.
At Stonewall Bank, currents were dominated by a strong southwestward (>0.5 m s–1) internal hydraulic flow (Moum and Nash 2000; Nash and Moum 2001). This flow produced interfacial shear instabilities between a plunging lower layer and the near-stagnant upper layer; intensified bottom boundary mixing and hydraulic jumps were also observed. Between the two experiments, a wide variety of T–S relations was observed at a range of turbulence intensities.
1) Patch selection
Single spectra of turbulence tend to be highly variable; in order to produce significant results, spectra must be ensemble-averaged to reduce the uncertainty and natural variability of the individual spectral estimates. Averaging is even more important to reduce the variability of composite spectra, which may rely on the difference between two spectral components of similar magnitudes. To increase the degrees of freedom of spectral estimates, the components Ψ
Turbulent patches were selected with regard to homogeneity of du′/dz, dC′/dz, and dT′/dz signals and uniformity of the mean gradient dS/dT. The T–S relation was required to be linear so that the relative contributions of T′ and S′ to C′ would remain constant. To illustrate how Ψ
Figure 3 illustrates the range of the spatial scales of the smallest velocity, temperature, and salinity fluctuations. For this patch, dC′/dz approximately represents dS′/dz, because the contribution of salinity gradient to conductivity gradient is much greater than that of temperature gradient. Note that the turbulent signal of dC′/dz in Fig. 3c is contained within the 20-cm band between 118.9 and 119.1 m; the lack of strong fluctuations outside this region indicates that the signal-to-noise ratio is high within the 20 cm layer.
2) Estimating scalar dissipation rates from spectra
If scalar spectra are fully resolved, then χθ is simply the complete integral of Eq. (6). In practice, measurements are limited by sensor response or noise at the smallest scales (or highest frequencies) and prevent complete integration of the scalar gradient spectrum. We define
A discussion of the frequency response of the microbead thermistor is given in Nash et al. (1999) and in the appendix (sec. a). The spatial response of the μC sensor is described in Nash and Moum (1999). For each of these sensors, corrections are applied in the frequency–wavenumber domain in order to restore lost variance. In addition, corrections were applied to account for the antialiasing filters and the imperfect response of the analog differentiators. Error and bias associated with the response corrections are discussed in the appendix (sec. a).

We remove the dependence of the theoretical shape on q by forming the nondimensional wavenumber αθ =
3. Observations
a. Temperature gradient spectra
To estimate χS, temperature fluctuations must be resolved in order to remove the contribution of Ψ
Spectra of temperature gradient closely follow the theoretical shape of Kraichnan (1968) especially near αT ∼ 1, the scales which contain most of the gradient variance. At the lower wavenumbers of the convective–diffusive subrange (αT ∼ 0.1), spectral amplitudes are significantly greater than those predicted by either the Batchelor or Kraichnan forms. Many investigators have observed a deviation in the convective–diffusive subrange, which may be attributed to remnant background vertical temperature structure. Dillon and Caldwell (1980) found that the deviation is greatest for small Cox numbers (
b. T–C cospectrum, coherence, and phase
The collapse of the normalized temperature gradient spectrum to Kraichnan's theoretical form gives us confidence that our temperature measurements are fully resolved. We will proceed to calculate the salinity gradient spectrum, which, using Eq. (9), depends on both Ψ
Case A: T′, C′, and S′ are in phase. The fluctuations in T′ and S′ are positively correlated on large scales, so that positive fluctuations occur simultaneously in both T′ and S′ and give rise to a positive fluctuation in C′. Even as T′ is attenuated at high wavenumbers (near k ∼
Case B: T′ is out of phase with both C′ and S′. This is the case where salinity dominates the conductivity signal on the overturning scale. Since T′ becomes attenuated at higher wavenumbers, S′ must also dominate C′ at the smallest scales. Hence, as long as T′ and S′ remain anticorrelated, T′ and C′ should also remain anticorrelated, as shown in Fig. 7 (case 𝗕). Note that the coherence is much lower for case 𝗕 than for case 𝗔. This is an indication that T′ and S′ are, in fact, decorrelating from each other at scales near ∼ 0.2
Case C: S′ is out of phase with both C′ and T′ on large scales. Conductivity is dominated by temperature on the energy-containing scales. However, above the thermal–diffusive wavenumbers (near ∼ 0.5
The cospectrum Ψ
c. Salinity gradient spectra and dissipation
In the previous sections, we identified and characterized the components of Ψ
The nondimensionalized spectrum of salinity gradient, shown in Fig. 8, approximately follows the universal form of Kraichnan. Only spectra with |Rρ| < 1 were used in this analysis (350 patches). These represent patches where the temperature contribution is less than 20 times the salinity contribution to the conductivity gradient spectrum. For |Rρ| > 1, the contribution of Ψ
In light of this, it is remarkable that the spectral estimates in Fig. 8 have such a narrow spread, given that a significant temperature contribution has been removed from Ψ
The dissipation of salinity variance χS is calculated by integrating Ψ
d. Covariance flux estimates
In Eq. (11), the angle brackets should ideally represent an average over the full spatial extent and temporal lifespan of a turbulent event. In practice, this is not possible from vertical-profiler measurements, and instead the averaging is performed with respect to a single dimension (z) instead of four (x, y, z; t). As a result of this undersampling, single-patch estimates of 〈w′θ′〉 are highly variable and may even be countergradient (Moum 1996a,b). Estimates of Fθ are thus only reliable when ensemble averaged over many turbulent events.
It was possible to unambiguously determine the background
4. Discussion
a. Universality of spectral shape
A comparison of the gradient spectra of T and S is shown in Fig. 11 as a summary. Two different normalizations illustrate the similarities and differences in the two spectral shapes. In Fig. 11a, the data are normalized using Kolmogorov scaling so that spectra collapse at the wavenumbers associated with the maximum variance in the velocity strain field (near kη). The scale separation between inertial subrange and diffusive (Batchelor) scales is 10 times greater for Ψ
In Fig. 11b, the data are normalized using Batchelor scaling so that spectra collapse in the viscous–diffusive subrange: Ψ
In the region where Ψ
In the viscous–convective subrange, spectral amplitudes are elevated over either universal shape so that the spectral slope in the viscous–convective subrange is less than +1. There are two possible reasons for this:
increased spectral intensity at low wavenumbers, as a result of the background vertical salinity structure [as has been suggested for T by Dillon and Caldwell (1980) and others] or
salinity variance within the viscous–convective subrange may return to larger wavenumbers in a reverse cascade, a result of incomplete mixing.
b. Flux comparisons
- based on the direct flux from covariance estimates,Fθwθ
- relating shear production to buoyancy production plus dissipation in the evolution equation of TKE,
The statistics of
Our estimates of
c. Differential diffusion by turbulence?
As a thought experiment to illustrate how differential diffusion might arise, consider the limiting case of mixing a scalar with infinitely small molecular diffusivity Dθ → 0, in a fluid parcel that evolves in time from quiescent to turbulent, and back to quiescent again. Assume that the scalar has some small effect on the buoyancy of the fluid. Scalar fluctuations produced by a turbulent overturn at large scales cascade to smaller scales as time evolves. After all turbulent motions subside, scalar gradients remain on a variety of scales but are not smeared by molecular diffusion. Given time, the scalar anomalies, each with a slight buoyancy anomaly, re-sort themselves and eventually return to the original scalar profile. The result? No irreversible mixing.



In the absence of persistent forcing,4 the lifespan of a turbulent patch may be considered to be τpatch ∼ O(N–1) (Crawford 1986; Moum 1996b). If τpatch is much shorter than the time it takes to cascade variance to the diffusive spatial scales, then there will be remnant salinity variance at moderately high wavenumbers (k >
The dependence of the flux ratios on Reb is shown in Fig. 15. Because of the small dynamic range of Reb and significant variability of our estimates, it is not possible to resolve a trend in our data. Also plotted for comparison is an estimated diffusivity ratio from Turner's (1968) laboratory experiments and the direct numerical simulations of Merryfield et al. (1998). Each curve exhibits a trend consistent with our intuitive arguments. Note that the DNS and lab experiments give quantitatively different ranges of Reb where KS/KT is significantly less than one.
In Turner's experiments, turbulence was generated by an oscillating grid that eroded a temperature or salinity interface. The rate at which fluid mixed across the interface was described in terms of an entrainment velocity (
To facilitate comparison with our oceanic data, we compute Reb for Turner's experiments from unpublished data.5 We estimate ϵ = u3/l based on the turbulent velocity (u = 8 × 10–6 × n m s–1) and length scale (l = 9 mm) at the interface, using the measurements of Thompson and Turner (1975), N2 = gΔρ/(ρΔz) was computed using Δz = 90 mm (the mean separation between interface and grid) in order to represent the background stratification. Numerically, Reb is approximately ∼ (90 − 270) Ri–1, where Ri is that in Turner (1973). The flux ratio was computed as d =
Also plotted is the ratio of cumulative salt flux ϕS to heat flux ϕT as estimated from the numerical simulations of Merryfield et al. (1998). Although Reb was not explicitly determined in their simulations, we estimate Reb as Fr2 Re (for runs I(a–c): red energy spectrum) and 12Fr2 (for runs II(a–c): blue energy spectrum). These indicate that 0.5 < KS/KT < 0.9 in the range 102 < Reb < 106, which is consistent with our data.
d. Relationship between spectral shape and KS/KT
Figures 5 and 8 show that the spectral shapes Ψ
We use geometrical arguments to determine the relationship between production and dissipation if the viscous–convective subrange scales as kn instead of k+1 as predicted by Corrsin (1951) in Eq. (5). First, we assume the wavenumber extent of the gradient spectrum is proportional to
Figure 16 shows the observed gradient spectra in relation to the Kraichnan universal spectrum, both in its original form and that with a k+0.85 viscous–convective subrange. We sort the data by buoyancy Reynolds number: the spectra in Fig. 16a are calculated from patches with Reb > 1000 and may be expected to be quasi-isotropic; the spectra in Fig. 16b are calculated from patches with low Reb where the influence of buoyancy may be significant. Clear differences in the spectra are observed at low wavenumber. For Reb > 1000, spectral levels of Ψ
We emphasize that we are unable to give a theoretical or analytical justification for our choice of spectral slope at this time: we simply suggest that a k+0.85 viscous–convective subrange is not inconsistent with our observed spectra (particularly at low Reb) and that such a slope followed by a diffusive cutoff ∝
5. Conclusions
Highly resolved measurements of salinity have been made from a vertical microstructure profiler near the coast of Oregon. Four hundred patches of turbulence were analyzed from which the spectrum of salinity gradient Ψ
We use do, the ratio of the transport of T and S on eddy scales, to compare the covariance fluxes of heat and salt, 〈w′T′〉 and 〈w′S′〉. Our finding that 〈do〉 ∼ 1 suggests that T and S are transported equally well by the large-scale eddy field. This contrasts our estimate of 〈dx〉 = 〈KS/KT〉, which describes the observed flux due to irreversible mixing and represents the ratio of eddy diffusivities. The distribution of dx exhibits much scatter (∼two decades), and has a geometric mean 〈dx〉 = 0.7 (Figs. 14 and 15). We present this value along with the following words of caution. As discussed in the appendix, there are several sources of error that may bias our estimates: we estimate that 0.6 < 〈dx〉 < 1.1 are probable bounds on our estimates of KS/KT. This error arises primarily from 1) our thermistor response corrections, 2) the assumption of isotropy in estimating χS and χT from one-dimensional vertical gradient spectra, and 3) our use of the Kraichnan universal spectrum (and the value of q), to correct for unresolved variance during spectral integration. In addition, these measurements represent a limited number of turbulent patches in a small subset of the oceanic parameter space; one important restriction within our analysis was the requirement that |Rρ| < 1. Furthermore, we have biased our data by selecting only turbulent patches for the analysis and neglecting weakly turbulent ones (for which differential diffusion is more probable). While our results are highly suggestive that dx < 1, our estimated uncertainty does not rule out the possibility that dx = 1, and from this limited dataset, we are unable to make a general claim for the value of KS/KT in the global ocean, or for its dependence on Reb.
Although the eddy motions produce variance of T and S proportional to (dT/dS)2, a value of dx < 1 would suggest that all of this variance is not dissipated. In particular, a disproportionate amount of the salinity variance is not being dissipated because of its low molecular diffusivity. It would be assumed that this variance eventually restratifies because the duration of the turbulent patch is not sufficient to allow complete irreversible mixing by molecular processes. This can only happen at low Reb and sufficiently high Ra, both of which imply weak, anisotropic turbulence. The observed spectral slope in the viscous–convective subrange for low Reb patches is consistent with an imbalance of production over dissipation.
While these are the first estimates of the irreversible salinity flux in the ocean, evidence that the irreversible transport for salt is less efficient than that for heat is not new. Our estimates of dx are consistent with those from the numerical experiments of Merryfield et al. (1998), who found the normalized flux ratio of salt to heat to be significantly less than one over a similar range of Reb, as shown in Fig. 15. The experiments of Turner (1968) suggest that differential diffusion (dx < 1) should be significant only at much smaller buoyancy Reynolds numbers (Reb < 100). The discrepancies between each of these experiments needs further investigation.
6. Potential consequences
The possibility that heat and salt are transported through the ocean at different rates has significant consequences. Most importantly, vast regions of the ocean are characterized by small buoyancy Reynolds number (Reb), where incomplete mixing is possible. Gargett and Holloway (1992) and others have suggested that small differences in the eddy diffusivities of heat and salt could have a significant impact on the thermohaline circulation.
Secondly, oceanographers have relied on the eddy diffusivity of heat—the quantity that the microstructure community usually measures—as being representative of that of salt, density, and buoyancy. While in regions of energetic turbulence this should be the case, in regions where salinity plays a dominant role in determining the stratification and where turbulence is weak it is likely that Kρ may be less than expected.
Finally, these results might further the interpretation of tracer-release experiments (Ledwell et al. 1993). The tracers used in those experiments generally have a molecular diffusivity comparable to that of salt,6 so that the inferred diffusivity is more closely related to KS than to KT or Kρ. It is therefore possible that these experiments tend to underestimate the true irreversible mixing of temperature or density, if indeed the turbulent diffusivities of heat and salt are unequal. This is especially likely in the regions of low turbulence levels where the dye release experiments have been relied upon to deliver bulk estimates of eddy diffusion coefficients. A quantitative understanding of the dependence of KS/KT on Reb will help to refine error estimates on diffusivities derived from such experiments.
These results highlight the need to develop a theoretical framework to relate the probability of incomplete mixing to flow parameters like the buoyancy Reynolds number. It should be possible to perform laboratory and numerical experiments to clarify some of the issues raised here and gain a more precise and fundamental understanding of dx. Future experiments should be designed to improve our understanding of the aspects that contribute to most of the uncertainty in the estimate of dx, namely the role of anisotropy in differential diffusion (and our ability to measure it) and the shape of the diffusive subrange of scalar gradient variance at a variety of Reb. More extensive measurements of both horizontal and vertical scalar gradient in the low-Reb regime will be needed to clearly determine the variability (or lack thereof) of KS/KT.
Acknowledgments
The authors have benefitted from informative discussions with Bill Smyth, Roland de Szoeke, Doug Caldwell, Jen MacKinnon, and Eric Kunze. The technical support of Michael Zelman, Mike Neeley-Brown, Ray Kreth, Gunnar Gunderson, and Greig Thompson made these measurements possible. We thank Stewart Turner for providing us with data from his 1966 experiments. This manuscript has benefitted from the critical and insightful comments of an anonymous reviewer. Funding by the Office of Naval Research and the National Science Foundation made this work possible.
REFERENCES
Altman, D. B., and A. E. Gargett, 1990: Differential property transport due to incomplete mixing in a stratified fluid. Stratified Flows, E. List and G. Jirka, Eds., American Society of Civil Engineers, 454–460.
Batchelor, G. K., 1959: Small-scale variation of convected quantities like temperature in turbulent fluid. J. Fluid Mech., 5 , 113–133.
Bogucki, D., J. Domaradzki, and P. Yeung, 1997: Direct numerical simulations of passive scalars with Pr > 1 advected by turbulent flow. J. Fluid Mech., 343 , 111–130.
Corrsin, S., 1951: On the spectrum of isotropic temperature fluctuations in isotropic turbulence. J. Appl. Phys., 22 , 469–473.
Crawford, W. R., 1986: A comparison of length scales and decay times of turbulence in stably stratified flows. J. Phys. Oceanogr., 16 , 1847–1854.
Dillon, T. M., and D. R. Caldwell, 1980: The Batchelor spectrum and dissipation in the upper ocean. J. Geophys. Res., 85 ((C4),) 1910–1916.
Gargett, A. E., 2002: Differential diffusion: An oceanographic primer. Progress in Oceanography, Pergamon, in press.
Gargett, A. E., and G. Holloway, 1992: Sensitivity of the GFDL ocean model to different diffusivities of heat and salt. J. Phys. Oceanogr., 22 , 1158–1177.
Gargett, A. E., and J. N. Moum, 1995: Mixing efficiencies in turbulent tidal fronts: Results from direct and indirect measurements of density flux. J. Phys. Oceanogr., 25 , 2583–2608.
Gargett, A. E., T. Osborn, and P. Nasmyth, 1984: Local isotropy and the decay of turbulence in a stratified fluid. J. Fluid Mech., 144 , 231–280.
Gibson, C. H., and W. H. Schwarz, 1963: The universal equilibrium spectra of turbulent velocity and scalar fields. J. Fluid Mech., 16 , 365–384.
Gregg, M. C., 1984: Entropy generation in the ocean by small-scale mixing. J. Phys. Oceanogr., 14 , 688–711.
Gregg, M. C., . 1987: Diapycnal mixing in the thermocline: A review. J. Geophys. Res., 92 ((C5),) 5249–5286.
Gregg, M. C., and T. B. Meagher, 1980: The dynamic response of glass-rod thermistors. J. Geophys. Res., 85 , 2779–2786.
Gregg, M. C., E. D'Asaro, T. Shay, and N. Larson, 1986: Observations of persistent mixing and near-inertial internal waves. J. Phys. Oceanogr., 16 , 856–885.
Head, M. J., 1983: The use of miniature four-electrode conductivity probes for high resolution measurement of turbulent density or temperature variations in salt-stratified water flows. Ph.D. thesis, University of California at San Diego, 211 pp.
Holloway, G., A. Gargett, and B. Merryfield, 2001: Differential mixing in bistable conditions. Proc. 'Aha Huliko'a Hawaiian Winter Workshop, Manoa, HI, University of Hawaii, 187–190.
Itsweire, E., J. Koseff, D. Briggs, and J. Ferziger, 1993: Turbulence in stratified shear flows: Implications for interpreting shear-induced mixing in the ocean. J. Phys. Oceanogr., 23 , 1508–1522.
King, D. B., and E. S. Saltzman, 1995: Measurement of the diffusion coefficient of sulfur hexafluoride in water. J. Geophys. Res., 100 ((C4),) 7083–7088.
Kolmogorov, A. N., 1941: The local structure of turbulence in an incompressible viscous fluid for very large reynolds numbers. Dokl. Akad. Nauk., 30 , 299–303.
Kraichnan, R., 1968: Small-scale structure of a scalar field convected by turbulence. Phys. Fluids, 11 , 945–953.
Ledwell, J. R., A. J. Watson, and C. S. Law, 1993: Evidence for slow mixing across the pycnocline from an open-ocean tracer-release experiment. Nature, 364 , 701–703.
Luketina, D. A., and J. Imberger, 2001: Determining turbulent kinetic energy dissipation from batchelor curve fitting. J. Atmos. Oceanic Technol., 18 , 100–113.
Merryfield, W. J., G. Holloway, and A. E. Gargett, 1998: Differential vertical transport of heat and salt by weak stratified turbulence. Geophys. Res. Lett., 25 , 2773–2776.
Moum, J. N., 1990: Profiler measurements of vertical velocity microstructure in the ocean. J. Atmos. Oceanic Technol., 7 , 323–333.
Moum, J. N., . 1996a: Efficiency of mixing in the main thermocline. J. Geophys. Res., 101 ((C5),) 12057–12069.
Moum, J. N., . 1996b: Energy-containing scales of turbulence in the ocean thermocline. J. Geophys. Res., 101 ((C6),) 14095–14109.
Moum, J. N., and J. D. Nash, 2000: Topographically induced drag and mixing at a small bank on the continental shelf. J. Phys. Oceanogr., 30 , 2049–2054.
Moum, J. N., M. C. Gregg, R. C. Lien, and M. E. Carr, 1995: Comparison of turbulence kinetic energy dissipation rate estimates from two ocean microstructure profilers. J. Atmos. Oceanic Technol., 12 , 346–366.
Nash, J. D., and J. N. Moum, 1999: Estimating salinity variance dissipation rate from conductivity microstructure measurements. J. Atmos. Oceanic Technol., 16 , 263–274.
Nash, J. D., . 2001: Internal hydraulic flows on the continental shelf: High drag states over a small bank. J. Geophys. Res., 106 ((C3),) 4593–4611.
Nash, J. D., D. R. Caldwell, M. J. Zelman, and J. N. Moum, 1999: A thermocouple probe for high speed temperature measurement in the ocean. J. Atmos. Oceanic Technol., 16 , 1474–1482.
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 , 256–271.
Osborn, T. R., and C. S. Cox, 1972: Oceanic fine structure. Geophys. Fluid Dyn., 3 , 321–345.
Schmitt, R. W., 1979: Flux measurements on salt fingers at an interface. J. Mar. Res., 37 , 419–436.
Smyth, W. D., 1999: Dissipation-range geometry and scalar spectra in sheared stratified turbulence. J. Fluid Mech., 401 , 209–242.
Smyth, W. D., and J. N. Moum, 2000: Anisotropy of turbulence in stably stratified mixing layers. Phys. Fluids, 12 ((6),) 1343–1362.
Sreenivasan, K. R., 1996: The passive scalar spectrum and the Obukhov–Corrsin constant. Phys. Fluids, 8 ((1),) 189–196.
Thompson, S., and J. Turner, 1975: Mixing across an interface due to grid turbulence generated by an oscillating grid. J. Fluid Mech., 67 ((2),) 349–368.
Turner, J., 1968: The influence of molecular diffusivity on turbulent entrainment across a density interface. J. Fluid Mech., 33 , 639–656.
Turner, J., . 1973: Buoyancy Effects in Fluids. Cambridge University Press, 368 pp.
Washburn, L., T. F. Duda, and D. C. Jacobs, 1996: Interpreting conductivity microstructure: Estimating the temperature variance dissipation rate. J. Atmos. Oceanic Technol., 13 , 1166–1188.
APPENDIX
Sources of Error and Bias
To understand the limits and significance of the Osborn–Cox estimate of diffusivity and associated fluxes, quantification of the error and/or bias introduced into the spectra (Ψ
Six significant sources of error that influence the preceding calculations are investigated and are summarized below.
(a) Signal attenuation: A careful description of the frequency response of the microbead thermistor has been determined by in situ comparisons with benchmark sensors. A theoretical and laboratory-verified spatial correction was applied to the conductivity probe. Such corrections to the power response can be determined to within a ±10% accuracy; since these corrections represent an average of 40% of the measured variance, the effect on the total measured variance is <10%.
(b) Sensor noise: To calculate χθ, gradient spectra have been integrated over a wavenumber range limited by
, the wavenumber where spectral amplitudes intersect an empirical noise spectrum. As a result, some noise may be included in high-k spectral estimates, causing the highest wavenumbers to be biased slightly high. Since spectral levels near the high wavenumbers are small, the effect of this noise on the total gradient variance (χS or χT) is small (0.6%).(c) Estimation of dS/dz and dT/dz: There are several ways that the mean gradients can be calculated; the average error is 3%.
(d) Cospectrum Ψ
: A lag between T and C results in aliasing variance from the cospectrum into the imaginary part of the cross spectrum. Such phase errors may have biased χS low by 2%.CzTz (e) Effect of anisotropy on χθ estimates from vertical spectra: Vertical gradient scalar spectra are enhanced over horizontal spectra at low Reb. Because motions become more isotropic at higher wavenumbers, the variance of salinity gradient is more isotropic than that of temperature gradient, which occurs at larger scales. As a result, estimates of χT may be biased high whereas estimates of χS are likely to be unbiased (or at least less biased). Estimates of χS/χT may be significantly biased low at very low Reb. Using an empirical relation derived from the direct numerical simulations of Smyth and Moum (2000), we find that χS/χT may be biased 10% low on average; however, the bias is less than 2% for Reb > 200.
(f) Choice of the form of the universal spectrum (Kraichnan or Batchelor), and the value of the constant q: This parameter relates the least principal strain rate to the dissipation rate (γ = −q–1
) and alters the wavenumber extent of the universal form for a given kb. The error that can be introduced into the ratio χS/χT by assuming an improper spectral shape for integration correction can exceed 20%. Any bias in our estimate of kb ∝ ϵ1/4 is implicitly considered (as a bias in q).
Given the magnitude of these individual sources of error, we can place error bounds on our estimate of dx = KS/KT. Thermistor attenuation, sensor noise, and error in estimation of mean gradients are random and add to produce an error of ±14%. Phase errors and the effect of anisotropy may have biased dx low by as much as −12%. The shape of the universal spectrum (and the value of q) used for integration correction has significant effect on dx; for plausible values of q, 0.68 < dx < 0.84. Combining each of these errors gives a maximum range for dx = KS/KT of 0.57 < dx < 1.06. For the purpose of discussion in this paper, we round this to 0.6 < dx < 1.1. From this error analysis, it is suggestive that dx < 1; at the same time, however, it is impossible to distinguish dx from unity, and thus KS = KT is possible.
Probe response
Temperature
During the Heceta Bank experiment, there were many turbulent patches where the conductivity was dominated by temperature, and Ψ
No turbulent patches with dT/dS > 20 K/psu were observed at Stonewall Bank, making it difficult to use the μC sensor to determine the thermistor time constant. Instead, the thermocouple was used as a benchmark, which we believe to respond without attenuation at 100 Hz (Nash et al. 1999).
Conductivity
Error from sensor noise
To avoid contamination of χT by sensor noise, gradient spectra are integrated over the subrange 0 < k <
In 96% of the patches, the contribution of sensor noise to the temperature gradient variance is less than 2%. Only in four of the weakly turbulent patches, where χT < 5 × 10–10 K2 s–1 does
Error in dT/dz, dS/dz
Estimates of Kθ are sensitive to error introduced in the determination of the background vertical scalar gradient dθ/dz. For example, the comparison of KS/KT is sensitive to error or bias in the ratio [(dT/dz)/(dS/dz)]2.
Errors specific to χS: The T–C phase
Since Ψ
While the phase lag ϕμC associated with the μC sensor is likely small (but remains to be quantified) over the low wavenumbers where Ψ
Because the μC phase response and the spatial separation between sensing volumes are not easily measured, we determined ϕlag = ϕμC + ϕFP07 + ϕsep empirically by assuming the linearized form of τFP07 such that ϕlag ≃ 2πfτlag. In the time domain, τlag simply represents a time lag between sensors and was chosen to produce zero phase difference between T and μC for the “average” temperature-dominated patch. For our μCT sensors, τlag = 14–18 ms.
Some deviation in the T–C phase from the expected 0° or 180° was observed from patch to patch.A1 To investigate the role that an improper phase lag would play on the estimation of χS, an analysis was performed that restricted the phase between C and T to either 0° or 180°. It was found that the mean difference between χS calculated in this manner and χS calculated from the observed phase was less than 2%. This small error can be rationalized by realizing that the cospectrum is sensitive to the cosine of the phase, so a 10° mismatch in phase (near ϕ = 0° or ϕ = 180°) only lowers the spectrum by 1.5%.
Bias associated with anisotropy
The ratio of the length scale where buoyancy effects are important [the Ozmidov scale: Lo = (ϵ/N3)1/2] to the length scale where viscous effects are important [the Kolmogorov scale: Lk = (ν3/ϵ)1/4] is Lo/Lk =
Vertical gradients are enhanced over horizontal gradients at scales O(Lo). At low -Reb, estimates of χθ based on Ψ
Because we have no way to directly measure the anisotropy of our turbulent patches, we quantify the effect of anisotropy using the direct numerical simulations of Smyth and Moum (2000, their Fig. 20c), which give a qualitatively similar relationship to that of Itsweire et al. (1993). Defining mθ =
Figure A6 indicates that
This Prandtl number dependent bias in estimating χ highlights the difficulty of understanding the Reb dependence of dx = KS/KT. This is because the Reb range where one might expect dx < 1 is precisely the same range where anisotropy affects our estimates of χT and χS. While the anisotropy of the flow at low Reb is a possible mechanism for generating unequal eddy diffusivities of heat and salt, at the same time it complicates our ability to measure those diffusivities. A more complete description of how anisotropy affects the diffusive scales in weakly turbulent, multicomponent flows seems necessary if we are to understand how KS/KT varies under a variety of Reb. Obtaining horizontal spectra may play a key role in this problem.
Error estimating χθ
The use of a universal scalar spectrum represents the largest source of error in estimating χθ from under-resolved spectra. The assumption of either the Kraichnan or Batchelor universal forms and the value of the parameter q affects the amount of variance assumed to be outside our limits of integration when determining χ from Ψ. Figure A7 shows the distribution of q as determined by comparing Ψ
While the bootstrap confidence limits on the mean are relatively small, the distribution of estimates spans a factor of 2. Only 50% of the estimates fall in the range 4.2 < qb < 7.5 and 5.5 < qk < 10.3. We use these as ranges to test the effect of the value of q on our estimates of χT and χS by using a universal spectral form for integration correction.
The effect of the value of q and the choice of the universal spectral form (Kraichnan or Batchelor) on the fraction of variance resolved by our measurements (i.e., the correction factor which needs to be applied to unresolved spectra) is shown in Fig. A8. The degree to which this affects the estimate of KS/KT can be significant; compared to estimates made using the Kraichnan form and qk = 7.5, a bias of 15% in dx could be introduced by choosing the incorrect universal form (Fig. A9). Note that the choice of q has less effect on KS/KT when the Kraichnan spectrum is used; this is likely the result of the Kraichnan spectrum more closely matching the observations. Since the value qk = 7.5 was chosen for this analysis because it best matches the spectral shape of scalar gradients, it would be surprising if the extreme values presented in Fig. A8 described the real spectra of our observations.
We also note that observed spectral levels of T and S in the viscous–convective subrange exceed those of theoretical forms (Fig. 11b). As a result, the use of Eq. (10) to correct for lost variance will overestimate χθ for poorly resolved spectra. Since T is generally better resolved than S (Fig. A8), χS/χT may be biased slightly high from this effect. However, the similarity of the nondimensionalized spectra of Ψ
Components of the salinity gradient spectrum Ψ
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
A side view of the upper inch of the microconductivity–temperature (μCT) probe (left, 2 × mag). The fast-response FP07 microbead thermistor (T) is separated by 1 mm from the conductivity (μC) tip, a cross section of which is shown at right, magnified 100×. The conductivity probe consists of two current-supplying (I–, I+) and two voltage-measuring spherical platinum electrodes (V–, V+) supported by a fused glass matrix (G). The sensor averages conductivity over a bipolar volume of radial extent ∼3 mm and has a −3 dB power attenuation near k ∼ 300 cpm. (Photographs courtesy Mike Head, Precision Measurement Engineering)
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
A 25-cm vertical segment of (a) du′/dz, (b) dT′/dz, and (c) dC′/dz within a turbulent patch at depth 120 m near Heceta Bank. The spatial scales of conductivity gradient are dominated by the salinity gradient and are ∼10 times smaller than those of temperature gradient, as indicated by the number of zero crossings in a given spatial interval, and is consistent with
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Gradient spectra of velocity, temperature, and salinity associated with a turbulent patch 3-m-thick (the data shown in Fig. 3 is contained within this patch). Smooth curves represent the universal forms of Nasmyth (for shear spectra) and Kraichnan (for scalar gradient spectra). The upper-left panel shows the two orthogonal components of velocity shear Ψ
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Summary of 407 nondimensionalized spectra of temperature gradient as a function of temperature-normalized wavenumber αT =
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
An aid to understanding T–C phase and coherence. Indicated on a T–S diagram are lines of constant conductivity (dotted) and three distinct regions (shaded) with T–S slopes that give rise to different T–C phase. The segments of the circle represent regions in T–S space where turbulent fluctuations (T′, S′, and C′ relative to the origin, 33 psu, 9°C) could occupy. The large-scale turbulent fluctuations represented on this diagram are generally aligned with the slope of the local T–S relation and form a line segment passing through the origin. Each of the three regions (
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Shading indicates the distribution of coherence (top) and phase (bottom) between dT′/dz and dC′/dz; error bars represent 95% bootstrap confidence intervals. Patches have been averaged over the three different ranges of Rρ identified in Fig. 6 which characterize the distinct trends in phase and coherence described in the text. The phase has been plotted only for estimates where the coherence is significant; the average significance level is 0.15, and varies with patch length
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Summary of 350 nondimensionalized spectra of salinity gradient as a function of salinity-normalized wavenumber αS =
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Comparison of the direct estimate of salinity variance dissipation χS with its proxy formed from χT and the square of the mean T–S gradient (dS/dT)2
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Turbulent vertical velocity w′ (left) and the associated instantaneous turbulent fluxes, w′S′ and w′T′ (right panels) for the patch shown in Fig. 4. Also shown are S and T along with the associated resorted profiles. On average, positive w′ is associated with positive S′ and negative T′, leading to the downgradient fluxes of heat (〈w′T′〉 = −7.1 × 10–6 K m s–1) and salt (〈w′S′〉 = 2.6 × 10–6 psu m s–1)
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Collapse of salinity (▴) and temperature (○) gradient spectra in the viscous–convective (a) and viscous–diffusive (b) subranges is accomplished by appropriate normalization. In (a) both Ψ
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Statistics of Γd, the dissipation flux coefficient based on the irreversible mixing on diffusive scales. Estimates were made from 350 patches with |Rρ| < 1
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Statistics of Γo, the flux coefficient based on the large-eddy transport. Data represents 76 patches with |Rρ| < 1 and both 〈w′T′〉 and 〈w′S′〉 significant at the 95% level
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Statistics of the diffusivity ratio based on large-eddy transports do = (
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Variation of the flux ratios do (based on the largeeddy transport) and dx (based on scalar dissipations) with buoyancy Reynolds number Reb ≡ ϵ/(νN2). For comparison, the dashed line represents the normalized salt to heat flux ratio (ϕS/ϕT) computed from Merryfield et al. (1998) for direct numerical simulation of two-dimensional turbulence. The solid line represents the ratio of saline to thermal entrainment velocity from the laboratory experiments of Turner (1968). We estimate Reb for their experiments as discussed in the text
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Gradient spectra calculated from patches with (a) high Reb and (b) low Reb exhibit subtle but important differences in spectral shape, especially at low αη. More than 150 patches contribute to each range of Reb; normalization and symbols are the same as in Fig. 11a; error bars represent 95% bootstrap confidence limits. The dashed curves represent Kraichnan's universal scalar gradient spectrum; the solid lines represent a modified Kraichnan spectrum in which the viscous–convective subrange scales with k+0.85 instead of k+1. An approximate agreement between observations and the modified universal spectrum (particularly for low-Reb) suggests that the spectral slope in the viscous–convective subrange may be less than 1
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Fig. A1. The frequency response [H2(f)] of the FP07 thermistors. Each shaded region represents the 95% bootstrap confidence interval for the average over at least 10 patches. The dark shading represents the response of the thermistor used at Heceta Bank, calculated as H2 = b2Ψ
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Fig. A2. Relationship between resolved wavenumber and the fraction of variance resolved by the fast conductivity probe γμC (left). A histogram of γμC is shown to the right; an average of 60% of the variance is resolved
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Fig. A3. Contribution of thermistor noise to estimates of χT. Dotted lines in the scatterplot (left) indicate cases where noise represents 1% and 10% of the measured temperature variance. On average, noise represents 0.6% of the signal (histogram, right).
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Fig. A4. Distribution of fractional error ΔdT/dS in dT/dS which results from estimating the mean background gradient of dT/dz and dS/dz using a linear regression with 95% confidence and two methods of resorting
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Fig. A5. Approximate empirical description for the anisotropy ratio, mθ, derived from Fig. 20c of Smyth and Moum (2000)
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Fig. A6. Estimated effect of anisotropy on the scalar dissipation ratio from vertical gradients alone, based on the empirical relation in Fig. A5. mS/mT represents the fractional amount
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Fig. A7. Distribution of qb (left) and qk (right) from the comparison of Ψ
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Fig. A8. Fraction of the salinity and temperature gradient variance resolved. The upper plot shows the distribution of the resolved wavenumber for salinity
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Fig. A9. Effect of the form of the universal scalar spectrum (used for integration correction), on the ratio of KS/KT
Citation: Journal of Physical Oceanography 32, 8; 10.1175/1520-0485(2002)032<2312:MEOTSF>2.0.CO;2
Here
One-dimensional spectra are calculated by dividing the patches into ∼0.5 m, 50% overlapped segments that are Hanning windowed, Fourier transformed, and ensemble averaged. Frequencies are converted to wavenumber using Taylor's hypothesis: k = f/Wo.
This method is typically used to estimate ϵ from slowly profiling devices without shear probes if one assumes a fully resolved temperature gradient spectrum and constant value for q.
For this analysis we consider the turbulent events to be freely decaying patches. Hence, our scaling is inappropriate if the duration of the forcing is much greater than N–1, an example being the persistent mixing associated with near-inertial internal waves (Gregg et al. 1986).
We thank Stewart Turner for locating his 35 year old Cambridge laboratory notebook and kindly providing this data to us.
The molecular diffusivity of SF6, used by Ledwell et al. (1993), is D
;thThe expected values of T–C phase over wavenumbers where molecular processes are insignificant is either 0° or 180°, depending on the T–S relation.