1. Introduction
The turbulent state of the upper ocean is of paramount importance when trying to understand transfers that take place between the ocean and atmosphere. The air–water exchange of heat, momentum, and gases are controlled by turbulent fluxes within the marine boundary layer (Anis and Moum 1995; Kantha and Clayson 2003; Bogucki et al. 2010). To determine turbulent flux, the simultaneous values of turbulent kinetic energy (TKE) dissipation rate ɛ and the temperature variance dissipation rate χ within the same part of the water column are needed. When both ɛ and χ are known, their combination can be used to infer the vertical mixing efficiency as well as the local vertical density diffusivity (Oakey 1982). The simultaneous measurements of the pair (ɛ, χ) can be performed by employing two separate sensors: one to measure the flow shear and the other to obtain temperature gradients. However, it is also possible to obtain both ɛ and χ from measurements of only one quantity, the temperature gradient spectra (Dillon and Caldwell 1980), which in principle should be an easier experimental task. For a detailed in depth review of oceanic microstructure measurements see- Lueck et al. (2002).
Historically, in situ measurements of small-scale temperature fluctuations were initially carried out by a thin film resistance thermometer (Grant et al. 1968), a thermocouple (Nash et al. 1999), or most frequently with fast response thermistors (Lueck et al. 1977; Dillon and Caldwell 1980; Gregg and Maegher 1980; Hill 1987). Small-scale shear is usually recorded using airfoils, an approach pioneered by Osborn (1980). Such direct measurements of TKE ɛ with airfoils within the upper ocean are frequently employed in modern microstructure research (Anis and Moum 1995). Yet, this approach is more involved than using temperature gradient spectra to infer the temperature variance spectra, the dissipation rate χ, and the TKE dissipation rates (Dillon and Caldwell 1980; Oakey 1982; Ruddick et al. 2000). However, the latter approach can be employed only if a proper universal function describing the temperature spectra is available (Dillon and Caldwell 1980). Usually, the universal function is assumed to be that derived by Batchelor (1959).
The purpose of experiments reported in this paper is to demonstrate that in nonstratified, high Reynolds number flows the temperature gradient spectra are best described by the Kraichnan spectral form (Kraichnan 1968) rather than by the Batchelor form. This assertion has been known from analyzes of the direct numerical simulation (DNS) results of Bogucki et al. (1997) and recently also observed in the experiments of Sanchez et al. (2011).
The paper is organized as follows. Section 2 will include discussion of the relation between ɛ and the location of the temperature dissipation peak
2. Review of temperature spectra in water: Batchelor and Kraichnan forms
When investigating incompressible, homogenous, and isotropic flows, the TKE dissipation rate
More detailed small-scale temperature information can be obtained using spectral turbulence theories (Monin and Yaglom 1971). In particular, Batchelor (1959) was the first to deduce the spectrum for scalar characterized by a Prandtl number Pr ≫ 1. The Prandtl number is the ratio of the kinematic viscosity to the thermal diffusivity, Pr = ν/D. Since water temperature is characterized by a Prandtl number Pr ≫ 1, the Batchelor theory has often been applied to describe temperature fluctuations in water. Later Mjolsness (1975) supplied a theoretical scalar spectrum formula based on an alternate theory of Kraichnan (1968). The Kraichnan spectra were observed in a number of DNS simulations: for nonstratified homogenous and isotropic turbulent flows (Bogucki et al. 1997; Yeung et al. 2002, 2004) and for sheared and stratified turbulence (Smyth 1999). The Kraichnan scalar spectra were also observed in oceanic experiments performed by Nash and Moum (2002), and in the recent lake measurements of Sanchez et al. (2011). This evidence points to the Kraichnan spectrum and accompanying theory as the correct description of temperature spectra in water. Despite this evidence many, researchers routinely use the spectrum proposed by Batchelor (Stevens and Smith 2004; Luketina and Imberger 2001; Ruddick et al. 2000) which may lead to erroneous results when computing TKE dissipation rates from temperature dissipation spectra.
It should be noted that the theoretical results normally use wavenumber k expressed in units of radian/m while the experimental spectra are typically expressed as a function of the cyclical wavenumber




















A number of values for
Kraichnan (1968) estimated
While no single values for qB and qK emerge, the majority of the reported results are in the range between 3 and 5 for the former, and between 3.4 and 7.9 for the latter quantity. Throughout the rest of the paper fixed values of
3. Experimental procedure and results
Near surface turbulence measurements were carried out in the Texas A&M University-Corpus Christi (TAMUCC) freshwater swimming pool having dimensions of 25 m × 12 m × 3 m deep, over the duration of one hour. The data were collected at a depth of around 25 cm. The pool was thermally nonstratified with uniform mean rms temperature fluctuations in both vertical and horizontal directions of around 0.01°C over a 1-m distance. The mean water temperature during the experiment was about 21°C with corresponding Pr of around 7. On the day of the experiment the wind speed was ~5 m s−1, while the pool surface was populated with quasi-randomly distributed surface waves a few centimers in amplitude.




Experiment instrument sketch: ROV with attached VMP200.
Citation: Journal of Physical Oceanography 42, 10; 10.1175/JPO-D-11-0214.1
The data were collected over the course of 15 runs, each lasting about 1–3 min. The collected data are labeled as runs B through P. In the selected runs, to test for possible contamination of vibrations the electric motor was switched off and the ROV continued its motion propelled only by its inertia. We have not observed any significant noise in the thermistor data that could be related to the propelling motor. This is attributed to the relatively large inertia of the whole system (>200 kg) when propelled by a relatively weak electric motor (<25 W).
Only subsets of selected runs (C, E, I, and J)—see Fig. 2—were chosen to maximize the resolved temperature variance. These required finding segments with low horizontal velocity w of the ROV and the minimal rms of mean velocity (Δw/w) (see appendix A). The time series of the speed of the ROV for runs C, E, I, and J are presented in Fig. 3. The sub segments selected for analysis marked in Fig. 3 are as follows: 15–16 s for run C, 61.5–64.5 s for run E, 24–29 s for run I, and 113–116 s for run J. The measured flow Reynolds numbers, flow speed, temperature, and energy dissipation rates and uncertainties (i.e., Δw/w, Δχ/χ) for the selected runs are collected in Table 1. The Reynolds number is defined here as
Water temperature time series for the runs C, E, I, and J.
Citation: Journal of Physical Oceanography 42, 10; 10.1175/JPO-D-11-0214.1
Velocity time series for the runs C, E, I, and J.
Citation: Journal of Physical Oceanography 42, 10; 10.1175/JPO-D-11-0214.1
Parameters of different runs.
The FP07 thermistor thermal inertia and the VMP electronics limit the observed thermistor response. The VMP-200 carries out a very sophisticated thermistor signal processing which minimizes the electronics noise so that the measured temperature spectra are only limited by the thermistor inertia—see Lueck et al. (2002).
The thermistor thermal inertia is related to the rate of heat transfer from the outside fluid to the thermistor bead through the combined fluid boundary layer and the FP07’s glass coating (Lueck et al. 1977; Gregg and Maegher 1980; Hill 1987; Nash et al. 1999). In general, the response of the FP07 depends on fluid velocity—see Fig. 4—and varies between thermistors. In this experiment to correct the FP07 data for the thermal inertia, a single pole response function was used
FP07 thermistor single pole-correction function,
Citation: Journal of Physical Oceanography 42, 10; 10.1175/JPO-D-11-0214.1



Shows the measured one-dimensional temperature dissipation spectra
Citation: Journal of Physical Oceanography 42, 10; 10.1175/JPO-D-11-0214.1
To find the temperature dissipation peak location a polynomial curve was used to approximate
As seen in Fig. 5, at low wavenumbers both spectral forms agree fairly well with the experimental data. The differences appear at higher wavenumbers, with the Kraichnan form describing the measured spectrum very well over the entire range of experimental wavenumbers. To see the match more clearly, the y axis value was changed to
Overall, the energy and temperature dissipation rates exhibit a relatively small scatter seen clearly in Table 1. This is attributed to the experimental conditions, specifically the fact that the measurements were taken as the ROV traversed horizontally in the swimming pool under essentially constant conditions including steady wind stress and the absence of any significant mean currents. This spatial homogeneity results in the dissipation spectra being measured under a background flow with a relatively constant Reynolds number, a situation not frequently experienced in oceanic conditions.
4. Parameters of the temperature variance spectra in the dissipation range
In the spectral forms proposed by Batchelor [Eq. (10)] and by Kraichnan [Eq. (11)] the TKE dissipation rate
Values of
The results of fitting the theoretical spectral forms to the measured data are presented in Fig. 6. The Kraichnan form is consistent with the observations over all experimental wavenumbers. In the case of the Batchelor form, the fit shows large systematic deviation from the measured data at large wavenumbers.
Measured temperature variance dissipation spectra and their best least squares fit to the Kraichnan and Batchelor spectral forms: (a) run C, 15–18 s; (b) run E, 61.5–64.5 s; (c) run I, 24–29 s; and (d) run J, 113–116 s.
Citation: Journal of Physical Oceanography 42, 10; 10.1175/JPO-D-11-0214.1
Figures 7 and 8 show the normalized temperature dissipation spectra in either linear–log or log–log scale. The data are clearly consistent with the Kraichnan spectral form and exhibits large deviations from the Batchelor form at high wavenumbers.
Normalized, compensated temperature spectra
Citation: Journal of Physical Oceanography 42, 10; 10.1175/JPO-D-11-0214.1
Normalized, compensated temperature spectra
Citation: Journal of Physical Oceanography 42, 10; 10.1175/JPO-D-11-0214.1
5. Conclusions
Temperature dissipation spectra were measured in a swimming pool using a fast thermistor FP07. In all analyzed cases the spectra were self-similar and followed closely the Kraichnan spectral form while displaying systematic deviations from the Batchelor form.
We found that the Kraichnan formula with the constant
As noted by Donzis et al. (2010) the constant
Based on the results in Table 1, for the kinetic energy dissipation rates, ɛ, obtained from the measured temperature dissipation spectra, the application of the Kraichnan functional form yields estimates of the kinetic energy dissipation rates by a factor of 2 to 3 smaller than the estimates inferred using the Batchelor form. We believe that the estimate based on the Kraichnan formula is more accurate.
Acknowledgments
DB acknowledges TAMUCC TRDF, NASA Physical Oceanography, and BP/The Gulf of Mexico Research Initiative support. We are grateful to Dr. Hua Li (JFE), Dr. John Gonzales (TAMUCC), and Mr. Chris Trombley (TAMUCC) for help with experiment.
APPENDIX A
The Temperature Dissipation Spectra
a. The location of the temperature dissipation spectra maximum 














b. Estimate of the measured temperature variance dissipation error
The “frozen field” hypothesis is applied to convert between spatial and temporal gradient:

APPENDIX B
Experimental Estimate of Thermistor Dynamic Response Function



In principle, the parameters of a thermistor transfer function can be inferred from experimental temperature spectra
To estimate the thermistor transfer function in the experiment two neighboring time intervals were selected such that the thermistor speed was significantly different between the measurements. The chosen intervals are at the beginning of Run J: from the 2nd to 4th second and from 5th to 7th second—see Fig. 3 with corresponding ROV speeds: w1 = 0.06 m s−1 and w2 = 0.14 m s−1.








The measured high wavenumber portion of function R(k) is shown in red. The blue lines denote the theoretical single-pole filter response when temperature dissipation is allowed to vary by a factor 0.3 between the two time segments. The corresponding double-filter response is marked by black lines. The undisturbed double pole filter function is represented by a black-dotted line for either double- or single-pole filter, respectively.
Citation: Journal of Physical Oceanography 42, 10; 10.1175/JPO-D-11-0214.1
The results show that the high wavenumber part (
By fitting the modeled R(
REFERENCES
Anis, A., and J. N. Moum, 1995: Surface wave–turbulence interactions: Scaling ε(z) near the sea surface. J. Phys. Oceanogr., 25, 2025–2045.
Antonia, R. A., and P. Orlandi, 2003: On the Batchelor constant in decaying isotropic turbulence. Phys. Fluids, 15, 2084–2086.
Batchelor, G. K., 1959: Small scale variation of convected quantities like temperature in a turbulent fluid. J. Fluid Mech., 5, 113–133.
Bogucki, D., A. J. Domaradzki, and P. K. Yeung, 1997: Direct numerical simulations of passive scalars with Pr > 1 advected by turbulent flow. J. Fluid Mech., 343, 111–130.
Bogucki, D., M.-E. Carr, W. Drennan, P. Woiceshyn, M. Schmeltz, and T. Hara, 2010: Estimates of gas transfer using satellite scatterometer during the 2001 GasEx experiment. Int. J. Remote Sens., 31, 75–92.
Dillon, T. M., and D. R. Caldwell, 1980: The Batchelor spectrum and dissipation in the upper ocean. J. Geophys. Res., 85, 1910–1916.
Donzis, D. A., K. R. Sreenivasan, and P. K. Yeung, 2010: The Batchelor spectrum for mixing of passive scalars in isotropic turbulence. Flow Turbul. Combust., 85, 549–566.
Gibson, C. H., 1968: Fine structure of scalar fields mixed by turbulence. II. Spectral theory. Phys. Fluids, 11, 2316–2327.
Grant, H. L., B. A. Hughes, W. M. Vogel, and A. Moilliet, 1968: Some observations of the occurrence of turbulence in and above the thermocline. J. Fluid Mech., 33, 443–448.
Gregg, M. C., 1999: Uncertainties and limitations in measuring ɛ and χT. J. Atmos. Oceanic Technol., 16, 1483–1490.
Gregg, M. C., and T. B. Maegher, 1980: The dynamic response of glass rod thermistor. J. Geophys. Res., 85 (C5), 2779–2786.
Hill, K. D., 1987: Observations on the velocity scaling of thermistor dynamic response functions. Rev. Sci. Instrum., 58, 1235–1238.
Kantha, L. H., and C. A. Clayson, 2003: Ocean Mixed Layer in Encyclopedia of Atmospheric Sciences. Academic Press, 2779 pp.
Kraichnan, R., 1968: Small-scale structure of a scalar field convected by turbulence. Phys. Fluids, 11, 945–953.
Lueck, R. G., O. Hertzman, and T. R. Osborn, 1977: The spectral response of thermistors. Deep-Sea Res., 24, 951–970.
Lueck, R. G., F. Wolk, and H. Yamazaki, 2002: Oceanic velocity microstructure measurements in the 20th century. J. Oceanogr., 58, 153–174.
Luketina, D. A., and J. Imberger, 2001: Determining turbulent kinetic energy dissipation from Batchelor curve fitting. J. Atmos. Oceanic Technol., 18, 100–113.
Mjolsness, R. C., 1975: Diffusion of a passive scalar at large Prandtl number according to the abridged Lagrangian interaction theory. Phys. Fluids, 18, 1393–1394.
Monin, A. S., and A. M. Yaglom, 1971: Statistical Fluid Mechanics: Volume 2. The MIT Press, 886 pp.
Nash, J. D., and J. N. Moum, 2002: Microstructure estimates of turbulent salinity flux and the dissipation spectrum of salinity. J. Phys. Oceanogr., 32, 2313–2333.
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.
Newman, G. R., and J. R. Herring, 1979: A test field study of a passive scalar in isotropic turbulence. J. Fluid Mech., 94, 163–194.
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., 1980: Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr., 10, 83–89.
Ruddick, B., A. Anis, and K. Thompson, 2000: Maximum likelihood spectral fitting: The Batchelor spectrum. J. Atmos. Oceanic Technol., 17, 1541–1555.
Sanchez, X., E. Roget, J. Planella, and F. Forcat, 2011: Small-scale spectrum of a scalar field in water: The Batchelor and Kraichnan models. J. Phys. Oceanogr., 41, 2155–2167.
Smyth, W. D., 1999: Dissipation-range geometry and scalar spectra in sheared stratified turbulence. J. Fluid Mech., 401, 209–242.
Stevens, C. L., and M. J. Smith, 2004: Temperature microstructure beneath surface gravity waves. J. Atmos. Oceanic Technol., 21, 1747–1757.
Yeung, P. K., S. Xu, and K. R. Sreenivasan, 2002: Schmidt number effects on turbulent transport with uniform mean scalar gradient. Phys. Fluids, 14, 4178–4191.
Yeung, P. K., S. Xu, D. A. Donzis, and K. R. Sreenivasan, 2004: Simulations of three-dimensional turbulent mixing for Schmidt numbers of the order 1000. Flow Turbul. Combust., 72, 333–347.