ARTICLE IN PRESS
Deep-Sea Research I 53 (2006) 387–407
www.elsevier.com/locate/dsr
Sheared turbulence in a weakly stratified upper ocean
I.D. Lozovatskya,c,, E. Rogetb, H.J.S. Fernandoa, M. Figueroab,d, S. Shapovalovc
a
Arizona State University, Environmental Fluid Dynamics Program, MAE, Tempe, AZ 85287-9809, USA
b
University of Girona, Campus de Montilivi, E-17071 Girona, Catalonia, Spain
c
P.P. Shirshov Institute of Oceanology, Russian Academy of Sciences, Moscow 117997, Russia
d
CICESE, Department of Physical Oceanography, Ensenada, B.C., Mexico
Received 23 December 2004; received in revised form 30 September 2005; accepted 5 October 2005
Available online 19 December 2005
Abstract
Microstructure, ADCP and CTD profiles taken in the North Atlantic along 531N under moderate and high winds
showed that the median of log-normally distributed kinetic energy dissipation rate e within the upper mixing layer is
1.5 107 W/kg and the layer depth, on the average, is 40 m. Assuming that mixing efficiency g is a constant (g ¼ 0:2),
the following scaling is proposed for the normalized eddy diffusivity:
K^ b ¼ K b =ku z ¼ ð1 þ Ri=Ricr Þp Pr1
tr ,
where K b ¼ ge=N 2 , N2 is the squared buoyancy frequency, u the surface friction velocity, Ri the local Richardson number,
Prtr ¼ 1 þ Ri=Rib the turbulent Prandtl number, p ¼ 1 or 2/3, Ricr ¼ 0.1 and Rib ¼ 0.1 or 0.05. The power-law function
with p ¼ 1 relates the asymptotes of Kb(Ri) to the buoyancy scale LN ðe=N 3 Þ1=2 at RibRicr and to the shear scale
LSh ðe=Sh3 Þ1=2 at Ri5Ricr. If p ¼ 2=3, the lengthscale LR ¼ ðe=N 2 ShÞ1=2 replaces LN in spectra of ocean microstructure
than due to the influence of local shear. This mixing regime corresponds to intermediate Richardson numbers
(0.25oRio2).
Alternatively, if g is not a constant, but an increasing function of Ri for 0oRio1, then K^ b ðRiÞ shows a very weak
dependence on Ri for Rio0.25. Numerical experiments using a one-dimensional q2e model with different Kb
parameterizations indicate that the measured mixed-layer depth agrees well with model results when the diffusivity is
parameterized with an Ri-dependent g (the GISS model approach). The modeled dissipation profiles, however, resembled
microstructure measurements better if g is treated as a constant and the proposed formula for K^ b ðRiÞ is used.
r 2005 Elsevier Ltd. All rights reserved.
Keywords: Turbulence; Dissipation; Diffusivity; Mixing parameterization; Upper ocean; Modeling
1. Introduction
Drift currents and associated wind-induced mixing are important components of upper oceanic
Corresponding author. Arizona State University, Environ-
mental Fluid Dynamics Program, MAE, Tempe, AZ 85287-9809,
USA. Tel.: +1 480 965 5597; fax: +1 480 965 1384.
E-mail address:
[email protected] (I.D. Lozovatsky).
0967-0637/$ - see front matter r 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.dsr.2005.10.002
dynamics. Depending on the wind stress and
variability of buoyancy flux at the sea surface,
turbulent mixing in the upper layer can produce
vertical transports of kinetic energy, momentum
and heat down to the underlying pycnocline or
across the sea surface to the atmosphere. The upper
ocean in high and mid-latitudes is usually well
mixed by sustained moderate and high winds. With
the exception of relatively short periods of intense
ARTICLE IN PRESS
388
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
convective mixing, the residual mean stratification
of the upper layer is weakly stable with an averaged
2
squared buoyancy frequency N~ ¼ ðg=r0 Þqr̄=qz of
6 2
or less; here r0 is the reference
about 10 s
density, qr̄=qz the mean vertical density gradient (z
is positive downward), and g the gravity. Small
density gradients and large vertical shears Sh ¼
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ðqu=qzÞ2 þ ðqv=qzÞ2
in
the
upper
layer
(Sh2105–104s2) provide favorable conditions
for developing Kelvin–Helmholtz (K–H) instability,
and thus local turbulent mixing, when the Richardson number, Ri ¼ N 2 =Sh2 , falls below a critical
value Ricr. Traditionally such turbulence is parameterized in terms of exchange coefficients also
known as eddy viscosity KM and diffusivity Kb. The
exchange coefficients are assumed constant (say
KMo and Kbo, respectively) for Ri5Ricr, Ri dependent at intermediate Ri whilst decreasing to the
molecular values at RibRicr. For locally generated
turbulence, this variation of KM and Kb can be
modeled as
K M K Mo ð1 þ bRiÞm and K b K bo ð1 þ bb RiÞn ,
(1)
where m, n, b, and bb are constants. Some
commonly used variables are, m ¼ 1/2, b ¼ 10,
n ¼ 3/2, bb ¼ 10/3 (Munk and Anderson, 1948);
n ¼ 1, bb ¼ 5 (Monin and Yaglom, 1975); m ¼ 2,
n ¼ 3, b ¼ bb ¼ 5 (Pacanowski and Philander, 1981,
hereafter PP-81); m ¼ 1.5, n ¼ 2.5, b ¼ bb ¼ 5, for
Ri40.25 (Peters et al., 1988, hereafter PGT-88),
n ¼ 3/2, bb ¼ 10, (Pelegri and Csanady, 1994); and
m ¼ 1/2, n ¼ 3/2, b ¼ 10, bb ¼ 20 (Paka et al.,
1999). Based on the direct measurements of microstructure in the ocean, Soloviev et al. (2001) recently
reported KM ¼ KbK0(1Ri/Ricr) for RioRicr ¼
0.25, which signifies a very weak dependence
on Ri at Ri50.1. A similar result has been found
for atmospheric stratified flows (Monti et al.,
2002). On the other hand, PGT-88 showed an
extremely steep variation KbRi9.6 for RioRicr,
but obviously, on physical grounds, the diffusivities
are expected to asymptote to their non-stratified
values as Ri-0. The discrepancy between different results, however, is curious as it points to
the unreliability of existing scaling for Ri5Ricr.
We believe that the difficulties of N2 calculations
in quasi-homogeneous layers, the limited bandwidth of airfoil sensors that are used to measure
small-scale shear in highly turbulent zones and
insufficient averaging of Kb and Ri are major
contributors to the high uncertainty of Kb(Ri)
at small Ri (see Appendix B for details). In addition, possible deviation of mixing efficiency g
(at subcritical Ri) from the assumed constant
value g ¼ 0.2 (Oakey, 1982) used for Kb calculations is also an important contributor to the mystery (e.g., Strang and Fernando, 2001a, b; Smyth
et al., 2001; Canuto et al., 2001; Fringer and Street,
2003).
In this paper, new microstructure measurements of the dissipation rate e, thermohaline, and
velocity structure in the upper layer of North
Atlantic are presented to further support the notion
of relatively weak growth of Kb at Ri5Ricr 0:1
(Paka et al., 1999). Note that the dependence of
eddy viscosity KM on Ri is believed to be weaker
than of that of Kb, but the critical Richardson
number seems to play the same role in the
demarcation of weakly and strongly stratified
regimes for both cases. Recent atmospheric studies
(e.g., Monti et al., 2002) indicate that KM may even
grow at Ri4Ricr possibly because of the increase of
momentum transfer by internal waves in stably
stratified boundary layers (Lee et al., 2005); a
parameterization based on this result has produced
better predictions in mesoscale atmospheric models.
Direct measurements of vertical momentum flux in
the marine environment remain technically challenging (Moum, 1998), and thus understanding of the
role of momentum transfer in oceans has made only
limited progress.
The data to be discussed herein (Section 2)
were obtained during the 9th cruise of R./V.
Akademik Ioffe (P.P. Shirshov Institute of Oceanology, Russian Academy of Sciences) along 531N
in April 2001 (Tereschenkov et al., 2002), and
they have already been utilized to study the
response of mixed-layer depth to atmospheric
forcing (Lozovatsky et al., 2005). The present
work focuses on statistics of the dissipation rate e,
Ri, Kb, and the buoyancy Reynolds number Rb
(Section 3) and on the dependence of local Kb
on Ri within the upper boundary layer (Section 4).
Here, a new version of Kb(Ri) scaling is proposed and we also compared the processed data
with several widely used mixing parameterizations. The effects of an Ri-dependent g on the
behavior of Kb(Ri), which greatly impacts relevant
numerical modeling, are illustrated and discussed in
Sections 5 and 6. The main results are summarized
in Section 7.
ARTICLE IN PRESS
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
2. Observations
ADCP, CTD, and microstructure profiling measurements were conducted at hydrographic stations
and were taken roughly every 30 miles along 531N
from the Labrador coast to the shelf of Ireland in
late April 2001. Vertical profiles of horizontal
velocity components u and v were obtained via a
ship-mounted ADCP equipped with GPS navigation while a Neil Brown Mark III (NBIS) profiler
measured the temperature–salinity–density in conjunction with a microstructure profiler MSS
(Prandke et al., 2000) that measured small-scale
shear [and therefore e(z)]. Severe weather prevented
MSS casts at a number of stations. Technical details
of the MSS sensors, a brief description of the
processing of small-scale shear signal and the
calculation of e are given in Appendix A. Temperature, salinity, and density profiles were processed
according to the WOCE standards (Millard et al.,
1990). The conductivity channels of NBIS and MSS
were controlled with the water sampled by an
Autosal 8400B salinometer. Appendix B contains an
error analysis of the main variables (Kb, N2, Sh2,
and Ri). In the near-surface layer (the depth range
down to 12–20 m), ship-based ADCP measurements are noisy (e.g., Firing, 1988; Chereskin et al.,
1989) and thus, taking into account 8 m ADCP
vertical resolution, the depth range 0–16 m was
excluded from our analysis. Microstructure measurements at depths above 10–15 m are usually
considered contaminated by ship rolling (e.g., PGT88; Lombardo and Gregg, 1989; Paka et al., 1999),
and therefore we decided to exclude microstructure
data above zcw ¼ 16 m from the analysis. This
389
allows matching of ADCP and MSS data and to
focus on the inner part of the upper mixed layer that
is free of surface-wave influence. This choice of zcw
is further supported by the probability distribution
function of e given in Section 3.
Standard meteorological observations were conducted at all stations. Winter storms and series of
atmospheric cyclones that cross the Atlantic often in
early spring significantly influence upper-layer
dynamics. The mean wind speed during the entire
cruise was 10.7 m/s. Fig. 1 shows the friction
velocity u averaged over the time of measurements
at each station (about 2.5 h). The mean wind speed
at the stations was 8.7 m/s. Three periods of strong
stormy winds were encountered during the sailing
across the ocean. The direction of the sea-surface
buoyancy flux Jb was convection favorable at all
stations except two (Fig. 1), but wind-induced
turbulence, rather than convection, dominated
mixing in the upper boundary layer. This is
evidenced in Section 3 by comparing characteristic
dissipation rates with the estimates of mechanical
and buoyancy production of turbulence. The
calculations of u and Jb were made with the
Matlab Air-Sea toolbox (http://sea-mat.whoi.edu),
which employs bulk formulae for air–sea fluxes.
3. Turbulence statistics in the mixing layer
3.1. The mixing layer depth and averaged dissipation
As pointed out by Brainerd and Gregg (1995), the
depths of mixed (hD) and mixing (he) layers are the
same for developed, quasi-stationary turbulence.
For decaying, or fossil, turbulence, hD should exceed
Fig. 1. Friction velocity u at the sea surface at drift stations of the 9th cruise of R./V. Akademik Ioffe, April 17–May 1, 2001. The stations
marked by filled circles are shown in Fig. 2.
ARTICLE IN PRESS
390
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
he. These ideas apply if externally induced mixing is
evolved within an existing upper quasi-homogeneous layer (UQHL) of the ocean. When active
turbulence penetrates into a sharp pycnocline, he can
exceed hD because of incomplete mixing at the
pycnocline. In this paper, we have specified he as the
depth (below highly turbulent near-surface layer)
where e(z) rapidly decreases from about 107 to
108 W/kg. Note that this definition has a certain
amount of subjectivity. It is based on our experience
in visual analysis of e(z) profiles of the mixed layer
while employing a formal threshold level for e
(107 W/kg) to identify the bounds of the mixing
layer. The highly intermittent nature of dissipation
profiles, even after substantial averaging, makes it
difficult to objectively determine he. The mixing layer
depth calculated semi-objectively is, however, acceptable here, given that we are not focused on the
analysis of he itself, but interested in a heterogeneous
record of e samples devoid of near-surface effects
(waves) and the influence of pycnocline turbulence.
The mixed-layer depth hD, nevertheless, was identi-
-9
0
-7
σθ
-5
-9
-7
fied objectively by the algorithm of Kara et al.
(2000), using the threshold level for variations of
specific potential density dsy ¼ 0.02sy without any
vertical averaging or interpolation.
Several MSS profiles of sorted specific potential
density sy(z) and averaged kinetic energy dissipation rate e^ðzÞ together with hD and he are shown in
Fig. 2 to illustrate different regimes of mixing in
UQHL (1 m spaced samples of e(z) were smoothed
by a four-point running averaging). At St. 922
(53.121N, 50.801W), e.g., where turbulence is confined to a shallow mixed layer (hD ¼ 18 m) under
relatively low winds, hD he . The depths of mixed
and mixing layers are identical at St. 935 (53.481N,
39.781W), where active turbulence generated during
the 2nd storm (see Fig. 1) produced a deep
he ¼ hD ¼ 82 m. When the wind ceased after this
storm (St. 936; 53.401N, 38.451W), the dissipation
profile showed a mixing layer (he ¼ 50 m) that was
shallower than the well-mixed layer based on
density profiles (hD ¼ 67 m). Here, the turbulence
starts to decay in the lower part of the mixed layer,
log 10 ε, (W/kg)
-5
-9
-7
-5
-9
-7
-5
ε
10
20
30
Depth (m)
40
50
60
70
80
90
100
110
27.2
27.4
27.6 27.3
27.4
27.5 27.2
27.3
27.4 27.1
27.2
27.3
σθ
922
935
936
948
Fig. 2. Examples of the dissipation (dotted lines) and specific potential density (solid lines) profiles showing the depths of mixed (larger
arrows) and mixing (smaller arrows) layers at several stations. St. 922 (53.121N, 50.801W): turbulence is confined in a shallow mixed layer
under relatively low winds. St. 935 (53.481N, 39.781W): active mixing induced by the 2nd storm all over the UQHL. St. 936 (53.401N,
38.451W): decaying turbulence or turbulence penetration to a limited depth in a well-mixed layer after passage of the 2nd storm. St. 948
(52.511N, 28.701W): the development of a mixed boundary layer at the beginning of the 3rd storm due to high-level penetrating turbulence.
Positions of the stations along the transect and the associated values of u are shown in Fig. 1.
ARTICLE IN PRESS
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
and surface-generated turbulence penetrates to a
limited depth. Finally, St. 948 (52.511N, 28.701W)
gives an example of the mixing layer he ¼ 42 m
which is deeper than the mixed layer, hD ¼ 38 m.
Energetic turbulence at this station, which was
associated with the beginning of the 3rd storm, has
entrained fluids from the pycnocline to deepen the
homogeneous layer. Note at St. 948, he penetrates
deeper into the pycnocline than the mixed-layer
depth, and he at this station clearly indicates the
depth where e(z) within the stratified layer starts to
decrease. Fig. 2 illustrates various turbulence
regimes encountered, but at the majority of stations
hD was in general agreement with he (see Fig. 3)
because during the cruise, turbulence in the upper
layer was generated predominantly by active wind
forcing. Note that only those stations where MLD
exceeded 10 m were used in this analysis. A
relatively narrow scatter of samples in Fig. 3
suggests that most of our observations belong to
quasi-stationary conditions. Only at three stations,
the mixing he and mixed hD layer depths differed
substantially. This occurred at the beginning of the
2nd storm (St. 933), whence turbulence quickly
penetrated to a depth of about 50 m, but without
destroying the stratification completely given that
the wind forcing has acted only a short time.
Therefore, formally detected MLD is still shallow
(hD ¼ 15 m). When the storm developed, both he
and hD achieved 80 m at St. 935 (see Fig. 2).
Conversely, at Sts. 943 and 946, under calm wind
conditions, turbulence was confined to the upper
15–20 m, but the thickness of upper homogeneous
mixed layer, which was developed during the
previous storm, was still about 40 m.
391
The main contribution of wind mixing to the
mean level of dissipation within the mixing layer e~ ml
is evident from Fig. 4, where the dependencies of e~ ml
on the wind stress production ews, buoyancy flux Jb,
and a combination thereof esf are shown. We
estimated e~ml at each station as the integrated
dissipation in the depth range heozozcw divided
by the thickness of this layer. The Ekman scale
LE ¼ u =f was used to evaluate the possible integral
contribution of wind-stress-induced dissipation ews
to the measured e~ ml . Using the law of the wall
analogy, we specified ews as u3 =0:1kLE , where
k ¼ 0.4 is the von Karman constant. The factor
0.1 before LE is used to link integral turbulence scale
with the most energy-containing outer turbulent
scale L0 within the mixed layer (see, e.g., Monin and
Yaglom, 1975). The scaling for combined boundary-induced forcing (wind and convection) can be
given as esf ¼ u3 =0:1kLMO , where LMO ¼ u3 =J b is
the Monin-Obukhov scale. According to Fig. 4b,
the correlation between e~ml and Jb is weak and the
buoyancy production accounts for only 9% of the
mean level of dissipation within the mixing layer.
The wind stress ews correlates well with e~ ml (Fig. 4c)
and accounts, on the average, for more than 90% of
e~ml . The combined effect, esf , is responsible for 95%
of the mean dissipation e~ ml (Fig. 4a). There is a
lower correlation between esf and e~ml than between
ews and e~ml due to the destructive role of the lowcorrelated pair J b e~ ml .
3.2. Probability distribution functions
In Fig. 5, a histogram of the logarithm of
dissipation rate ðlog10 eÞ is shown for the depth
Fig. 3. The depth of the turbulent (mixing) layer he vs. the mixed-layer depth hD obtained from MSS profiles.
ARTICLE IN PRESS
392
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
Fig. 4. The dependences of the mean dissipation within the mixing layer on wind stress production ews (c), buoyancy flux Jb (b), and a
combined effect of the two esf (a).
Fig. 5. Histogram of the logarithm of kinetic energy dissipation
rate in the depth range zcwozohe (b); he is the depth of the upper
turbulent layer (mixing layer), zcw ¼ 16 m is the lower boundary
of the near-surface layer. Gaussian approximation is shown by
heavy line.
range zcwozohe (i.e., below the contaminated zone
from surface effects). The probability distribution of
e within the mixing layer at depths z4zcw is fairly
log-normal (Fig. 5) with m ¼ /log eS ¼ 15.73,
RMS(log e) ¼ 1.76 and small values of skewness
(0.07) and kurtosis (0.44), which must be zero for
normal distribution of natural logarithm of the
dissipation log e. The corresponding mean and
median values of e are /eS ¼ 3.5 107 and
med(e) ¼ 1.5 107 W/kg.
The
Kolmogorov–
Smirnov (K–S) goodness of the fit for normal
distribution of log e is 0.03, which exceeds a critical
value of K–S statistic at the 90% significance level
(0.028) while being almost equal to the K–S
goodness at the 95% level (0.031) of significance.
If log10 e probability distribution is calculated using
all 1 m averaged samples pertinent to the upper
turbulent layer, including those from the nearsurface layer (not shown here), then it substantially
departs from log-normal approximation with a long
tail of almost equal probability in the range
4.5olog10 eo2. This indicates that very large
values of e, which were mostly observed at zozcw,
belong to a different statistical population than
those having log10 ep4, formally justifying the
separation of near-surface and inner sections of the
mixing layer at zcw ¼ 16 m.
The turbulence in the upper layer is generated
mainly by wind stress, and therefore the generation
mechanism below the wave-breaking zone can be
construed as due to shear instability of drift
currents. As such, it is useful to analyze this layer
with the probability distribution function of the
gradient Richardson number Ri ¼ N 2 =Sh2 , where
N2 and Sh2 are the squared buoyancy frequency and
ADCP vertical shear, respectively. The buoyancy
frequency was computed with 4 m vertical resolution along monotonically arranged MSS density
profiles obtained by a Thorpe-sorted algorithm. The
shear was calculated with 8 m vertical separation
ARTICLE IN PRESS
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
(Appendix B) and interpolated to the depth of N2
profiles. The Sh2 and N2 records were low-pass
filtered with running averaging over four points and
then interpolated to Dz ¼ 1 m, allowing the calculation of eddy diffusivity K b ¼ ge=N 2 , where g is the
mixing efficiency traditionally taken as g ¼ 0.2
(Osborn, 1980; Oakey, 1982, 1985; Oakey and
Greenan, 2004). Many authors (see, e.g., recent
review of Peltier and Caulfield, 2003) suggest,
however, that in the range 0oRio1, g is a growing
function of Ri. In Section 5, we discuss how this
dependence can change the values of Kb and
therefore the Kb(Ri) dependence.
The cumulative distribution function of the
Richardson number F(Ri) given in Fig. 6 shows
that in the depth range zcwozohe the median of Ri
is as low as 0.1 and the probability of Rio0.25 is
above 60%. This distribution as well as that of e (see
Fig. 5) is well approximated by a log-normal
probability law. Note that we did not find any
noticeable correlation between the local Ri in the
depth range zcwozohe and the friction velocity u
at the sea surface for u p1:5 102 m=s. Only at
four stations, where u exceeded 1.8 102 m/s
(winds above 15 m/s), was a weak tendency noted
at lower Ri.
The state and intensity of mixing in UQHL can
be specified by two related variables Kb and the
393
buoyancy
Reynolds
number
Rb ¼ e=30nN 2
(Gibson, 1980; Stillinger et al., 1983). The latter
can also be expressed as Rb ¼ K b =6n signifying the
ratio between the eddy diffusivity and molecular
viscosity n. The distribution of buoyancy Reynolds
number F(Rb) in Fig. 6 shows a wide range, 10–106,
indicating highly energetic turbulence almost everywhere in the mixing layer. This turbulence can be
considered as locally isotropic if the criterion
ðe=nN 2 Þ3=4 4200 is satisfied (Gargett et al., 1984).
More than 95% of the observed Rb samples comply
with this threshold, Rb440 (see Fig. 6). The median
of Rb in Fig. 6 is as high as 4 103. The
corresponding log-normal approximation (the
straight dashed line) fits more than 80% of the
upper end of F(Rb) distribution starting from
Rb ¼ 6 102 . Perhaps, this signifies the lower
cutoff for fully developed isotropic turbulence
with cascade of turbulence energy, which is characterized by log-normal distribution of the dissipation (Gurvich and Yaglom, 1967). In the layers
with Rb4600, Kb exceeds the molecular viscosity
by a factor of more than 3500. About 15% of
samples at the lower end of F(Rb) depart from
log-normal approximation and possibly belong
to the base of UQHL, where internal-wave breaking generates patches of weak intermittent turbulence.
Fig. 6. The cumulative distribution functions of the Richardson F(Ri) and buoyancy Reynolds F(Rb) numbers in the depth range
zcwozohe. The corresponding log-normal approximations are given by straight and dashed lines. The median of Rb is 4 103; for Ri, the
median is 0.1, and the probability of Rio0.25 exceeds 60%.
ARTICLE IN PRESS
394
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
4. The diffusivity Kb calculated with a constant
mixing efficiency c
4.1. Vertical profiles
The importance of shear instability in generating
turbulent mixing within UQHL along the 531N
transect is illustrated in Fig. 7 using the vertical
profiles of diffusivity Kb and Ozmidov scale, LN ¼
ðe=N 3 Þ1=2 ¼ ðK b =gNÞ1=2 as well as sorted density sy,
shear, and the Ri at two typical Sts. 948 (developing
turbulence at the beginning of the 3rd storm) and
936 (active, but decaying turbulence after the 2nd
storm). The original profiles of the dissipation rate
e(z) are also shown in Fig. 7 to emphasize that the
general decrease of Kb and LN with depth is not a
simple consequence of a stratification change, and
also due to decreasing turbulence. It is interesting
that the choice of the mixing layer depth he made
semi-objectively from e(z) profiles roughly corresponds to the depth where Ri is close to 0.25 (St.
948) or even below 0.25 (St. 936). The averaged
depth of the mixing layer calculated over all
available e(z) profiles is /heS ¼ 48 m; the depth,
where Rio0.1–0.25, is 38 m on the average. At the
base of the mixing layer, Ri 1, the non-linear limit
for shear instabilities.
The laboratory stratified shear flow mixing data
of Strang and Fernando (2001b) suggest that
Rio0.1 corresponds to a regime of negligible
density stratification effects, 0.1oRio0.25 corresponds to K–H instability dominated regime and
0.25oRio1 is characterized by both internal-wave
breaking and K–H instabilities. Interestingly, at
Ri 1 the K–H instabilities and internal waves may
resonate, thus giving rise to the maximum mixing
efficiency (also see Strang and Fernando, 2001a;
Pardyjak et al., 2002). For Ri41, the flow is
dominated by Hölmböe instabilities, which have
weaker mixing behavior. It appears that the mixing
layer base is at a state of maximum mixing
efficiency, and a rise of Ri therein will lead to a
rapid decrease of shear-induced mixing. The results
also suggest that he is associated with the depth of a
critical Richardson number, which varies in the
range Ricr ¼ 0.1–1, depending on the specifics of
local generation processes.
The diffusivities in the mixing layer are high
(102–101 m2/s), but Kb sharply decreases to
105–104 m2/s when entering the pycnocline.
Slightly below the base of the mixing layer, at
depths where Ri 1, the Ozmidov lengthscale
decreases to about 1 m, which is a characteristic
lower cutoff of the inertial subrange of the kinetic
energy spectra of stratified turbulence in the oceanic
thermocline (Gargett et al., 1981). In all, local shear
instability appears to favor the generation of
turbulence in UQHL.
4.2. Kb(Ri) parameterizations
4.2.1. Critical Ri
Two Kb(Ri) dependences obtained at Sts. 933 and
939 are given in Fig. 8. Arguably, these examples
represent the typical quasi-stationary balance of
turbulent kinetic energy in the mixing layer below
the near-surface highly turbulent zone. The stations
were taken after the 1st and 2nd storm, exhibiting
relatively high dissipation in the depth range
24–48 m where Rio1. In both cases, Kb tends to
decrease with the increase of Ri. The trends can be
approximated as
Kb ¼
K Mo
Pr1 ,
ð1 þ Ri=Ricr Þp t
(2)
where Prt ¼ KMo/Kbo is the turbulent Prandtl
number, which in general is a function of Ri and
often written as
r
Pr1
t ¼ at ¼ ð1 þ Ri=Ri b Þ .
(3)
Here, KMo and Kbo are the eddy viscosity and
diffusivity in non-stratified flow, respectively, Ricr
and Rib are certain critical Richardson numbers,
which do not necessarily have the same numerical
values; Ricr is associated with transition from a nonturbulent to a turbulent regime (or from ‘‘weak’’ to
‘‘well-developed’’ turbulence) whereas Rib indicates
a specific state of stratified shear turbulent flow
when vertical mixing (buoyancy flux) starts to be
affected by stratification more significantly than the
momentum flux. For Ricr, the canonical values are
0.25 and 0.20 (e.g., Monin and Yaglom, 1975; PP81; PGT-88), although Ricr ¼ 0.1 has also been
suggested (e.g., Munk and Anderson, 1948; Lozovatsky et al., 1993; Pelegri and Csanady, 1994;
Lozovatsky et al., 2000). The data of Strang and
Fernando (2001a, b) and Pardyjak et al. (2002) as
well as numerical simulations of various turbulent
flows reviewed recently by Peltier and Caulfield
(2003), including DNS (Jacobitz et al., 1997)
indicate that for Ri40.1 the turbulence is affected
by buoyancy forces. The critical Rib has often been
set to Ricr (e.g., PP-81; PGT-88; and others), but
ARTICLE IN PRESS
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
395
Fig. 7. Profiles of the dissipation rate (1), diffusivity (2), and Ozmidov scale (3) (left) and the corresponding profiles of sorted density (4),
shear (5), and Ri (6) (right) at St. 936 (upper plot) and St. 948 (lower plot). The horizontal long and short dashed lines and the straight line
show the depths where Ri ¼ 0.1, 0.25, and 1, respectively. The crosses mark the mixing layer depths he, which is near the depth where
RicrE0.1/0.3. The Ozmidov scale decreases to 1 m at the depth, where RiE1 is observed.
ARTICLE IN PRESS
396
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
Fig. 8. The eddy diffusivity K b ¼ 0:2e=N 2 vs. the Richardson
number Ri at St. 933 (upper panel) and St. 939 (lower panel). The
symbols are 1 m averaged samples. The modeling dependencies
Kb(Ri) given by Eqs. (2) and (3) are shown by the bold (r ¼ 1,
p ¼ 2/3, Ricr ¼ 0.1, Rib ¼ 0.05) and thin (r ¼ 1, p ¼ 1,
Ricr ¼ Rib ¼ 0.1) continuous lines. The parameterization of
Pacanowski and Philander (1981) and Peters et al. (1988) are in
the thin and bold dashed line, respectively.
different values for Ricr and Rib have also been used
in numerical calculations and data analyses
(Ricr ¼ 0.1, Rib ¼ 0.3, Munk and Anderson, 1948;
Ricr ¼ 0.1, Rib ¼ 0.05, Paka et al., 1999; Toorman,
2000). More elaborate formulations of Kb and at
than those of Eqs. (2, 3) have been suggested, e.g.,
by Henderson-Sellers (1982), who analyzed the
measurements of Linden (1979) and Ueda et al.
(1981) and introduced KbKbo(1+(Ri/Ricr)2)1
with Ricr0.165 and at ¼ 1 þ 0:74Ri=1 þ 37Ri2 .
This leads to the lowest Rib 0:02, if the formula
for at is reduced to Eq. (3). Recent atmospheric
turbulent measurements of Monti et al. (2002) also
support Eq. (3) but require Rib ¼ 0.05 and r ¼ 1/2.
The majority of previous studies, however, prefer
r ¼ 1. This in turn yields a constant value for the
flux Richardson number Rf ¼ aRi at high Ri, which
is consistent with the assumption of a constant
mixing efficiency g (Oakey, 1982) when Kb is
calculated. Any r41 leads to a reduction of Rf at
high Ri, as evident from the atmospheric data of
Pardyjak et al. (2002).
4.2.2. Kb asymptotics and links to the turbulent
scales
As was stated in Section 1, empirical values of p
range between 0.5 (Munk and Anderson, 1948) and
2 (PP-81). If the eddy viscosity KMo in a nonstratified sheared ocean flow is specified using the
vertical shear Sh, and turbulent kinetic energy q2 or
the dissipation rate e as K Mo q2 =Sh, or K Mo e=Sh2
(e.g., PGT-88), then the natural choice for p would
be p ¼ 1, leading at RibRicr to K M q2 =N and
K M e=N 2 , respectively. These are the generally
accepted formulae for eddy viscosity in a stratified
ocean. The above asymptotic form for KM derived
using Eq. (2) is in consonance with the assumption
that, in a stratified layer without a local source of
shear energy production, the local turbulence scale
ltr should be equal to the Ozmidov scale, ltr ¼ LN,
thus yielding K M l tr q. Accordingly, in non-stratified shear flows, l tr ¼ LSh ðe=Sh3 Þ1=2 , where LSh is
the Tchen (1953) shear scale.
In an intermediate range of the Richardson
numbers (0.25oRio2), the scale ltr can be
specified as LR ¼ ðe=N 2 ShÞ1=2 (Lozovatsky et al.,
1993), leading to K M e=ðN 2 ShÞ2=3 , requiring p ¼ 2/3.
The lengthscale LR, which can also be written as
ul =N, ul being the local friction velocity, replaces
LN in the spectra of turbulent kinetic energy E(k), if
a production subrange with EðkÞðe=ShÞk1
emerges between the inertial EðkÞe2=3 k5=3 and
buoyancy EðkÞN 2 k3 subranges due to locally
generated shear-induced turbulence (Lozovatsky,
1996). Therefore, p ¼ 2=3 is a viable choice for
diffusivity parameterization, if turbulence is developed by a combination of non-local (wind stress in
our case) and local (inertial oscillations, for
example) shear sources.
For comparison, the exponent p ¼ 1/2 (Munk
and Anderson, 1948) gives K M e=ðNShÞ for
RibRicr, which does not correspond to any known
turbulent spectral lengthscale. Yet it correctly
reflects the dependence of KM on governing parameters of stationary stratified turbulence. Two other
putative values are p ¼ 2 (PP-81) and p ¼ 3/2 (e.g.,
PGT-88; Pelegri and Csanady, 1994), which produce K M eSh2 =N 4 and K M eSh=N 3 , respectively.
To satisfy stationary turbulent kinetic energy
budget, the flux Richardson number (or g), in these
cases should be a decreasing function of Ri, e.g.,
gRi1 for p ¼ 2 and gRi1/2 for p ¼ 3/2, which is
unlikely in weakly stratified UQHL. Note neither
PP-81 nor PGT-88 links K Mo with e=Sh2 , rather
they simply give a specific numerical value for the
ARTICLE IN PRESS
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
nominator of Eq. (2). Because Eq. (2) with p ¼ 2/3
and/or p ¼ 1 links Kb(Ri) to a specific spectral
structure of turbulent fluctuations (via corresponding lengthscales LR and LN), we favor these values
of p in Eq. (2) compared to traditional p ¼ 1/2, 3/2
and 2.
In all, r ¼ 1 is probably the most rational choice
for at(Ri) in Eq. (3). Also, Ricr and Rib were selected
from a set of values already discussed (Ricr ¼ 0.1 or
0.2, and Rib ¼ 0.05 or Ricr ¼ Rib ¼ 0.1 or 0.2). Eqs.
(2) and (3) fit individual data samples of Kb at St.
933 and St. 939 reasonably well with p ¼ 2/3 or 1
(Fig. 8). The scaling values of eddy viscosities KMo
are 5 102 and 4.5 102 m2/s for St. 933 and
4 102 and 3.5 102 m2/s for St. 939, respectively. While p ¼ 2/3 requires Ricr ¼ 0.1 and
Rib ¼ 0.05 for a best fit, the exponent p ¼ 1 (which
corresponds to the Ozmidov scale LN) fits the
data equally well with Ricr ¼ Rib ¼ 0.1. The parameterization of PP-81, with the original set of
p ¼ 2, r ¼ 1, Ricr ¼ Rib ¼ 0.2, but with a slightly
higher K Mo ¼ ð3 3:5Þ 102 m2 =s (compared to
102 m2/s), gives the best fit, but its standard error is
more than that of the first two approximations (see
the thin dashed lines in Fig. 8). To be able to fit the
data by using Eq. (2) with p ¼ 3/2 (e.g., Pelegri and
Csanady, 1994), it is necessary to vary the values of
Ricr and Rib for different stations, which is not
justifiable.
4.3. The normalized diffusivity
In order to compare
ities observed under
individual Kb samples
stations with different
and parameterize diffusivvarious wind conditions,
obtained at a number of
friction velocities u were
397
normalized using the Monin-Obukhov (1954) similarity theory, according to which the vertical
diffusivity in the surface boundary layer is given
by K sf ðzÞ ¼ ku z, where k is the Karman’s constant
and u the friction velocity. Based on Eq. (2), the
normalized diffusivity Kbn ¼ Kb/Ksf can be written
as
K bn ¼
Kb
1
.
¼
ku z ð1 þ Ri=Ricr Þp ð1 þ Ri=Rib Þ
(4)
Fig. 9a shows the bin-median values of K^ bn and
^ fitted by Eq. (4) for two sets of parameters p,
Ri
Ricr, and Rib. The bootstrap 90% confidence limit
for K^ bn was calculated for 1000 resampled points
when the actual number of samples was 15 for each
bin. The data were taken from Sts. 933, 936, 939,
and 948, where MLD ¼ 38–56 m and the wind
speed varied between 4.5 and 14.3 m/s. Both lines in
Fig. 9a fit the bin-median samples quite well in the
^ range from 1.3 102 to 5 101, where the
Ri
coefficients of determination are 0.80 and 0.77 for
^
lines 1 and 2, respectively. At very low Rio0:01,
the
data significantly depart from Eq. (4). This could be
attributed to insufficient accuracy of N2 calculations
when density gradients become very small. A
^ in Fig. 9a may,
decrease of K^ bn at the lowest Ri
^
however, signify a tendency to be independent of Ri
^
as Ri ! 0 because of vanishing buoyancy effects.
Also clear from Fig. 9a is that our data do not
support an explosive growth of the diffusivity at
^
Rio0:25
as reported by PGT-88 for patch turbulence in the equatorial thermocline.
Recently, Soloviev et al. (2001) parameterized the
normalized diffusivity K b =ku z as (1Ri/Ricr) for
RioRicr ¼ 0.25 and used the PGT-88 scaling for Kb
at Ri4Ricr ¼ 0.25. A slightly modified version of
Fig. 9. The bin-median estimates of the normalized diffusivity K^ b ¼ K b =kzu at the probability-equal Ri-intervals (large circles) with 90%
bootstrap confidence limits shown for both variables. The data are superimposed by the modeling functions given by Eq. (4) (left panel)
and by Eq. (5) (right panel).
ARTICLE IN PRESS
398
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
the Soloviev et al. (2001) original formula for the
normalized Kb is
Kb
K^ bS ¼
ku z
¼ 1 Ri=Ricr þ
K^ bS ¼
5 102
ku zð1 5RiÞ2:5
5 102
ku zð1 5RiÞ2:5
ðRi4Ricr Þ,
ðRioRicr Þ,
ð5Þ
where Ricr ¼ 0.25. The authors showed that the
results of microstructure measurements taken in the
upper turbulent layer of the western Pacific warm
pool are in general agreement with Eq. (5). Because
the foundation of Eq. (5) closely corresponds to that
of Eq. (4), it is instructive to compare the two
formulae with respect to the bin-median diffusivities
obtained in the upper layer of the North Atlantic.
Function (5) is presented in Fig. 9b for the same
experimental data set. If K^ bS is calculated using
Ricr ¼ 0.25, the result substantially departs from
^ but when Ricr
our measurements at intermediate Ri,
is reduced to 0.1, the agreement with experimental
data becomes much better (see Fig. 9b). This later fit
(using Ricr ¼ 0.1) is almost equally good as
that obtained with formula (4), which is shown in
Fig. 8a. The corresponding coefficients of determination are 0.7 and 0.74 for Eq. (4) and 0.6 for
Eq. (5), if the two samples at the highest and lowest
^ are disregarded. Note that Eq. (5) of Soloviev et
Ri
al. (2001) with Ricr ¼ 0.25 lies slightly above their
averaged data points (see Fig. 10 of Soloviev et al.,
2001) and this effect is manifested in our data also.
Fig. 10. Mixing efficiency g as a function of the Richardson
number Ri: (1) Strang and Fernando (2001a, b); (2) stable
nocturnal ABL; (3) Mellor and Yamada (1982); (4) Canuto et
al. (2001).
This suggests that the use of Eq. (5) with lower Ricr
(Ricr ¼ 0.1) should improve the performance of this
parameterization, thus making it comparable with
Eq. (4).
5. Evaluation of Kb(Ri) using Ri-dependent mixing
efficiency c
The analyses in Section 4 as well as the calculation of Kb and Rb (Section 3) rely on the hypothesis
of a constant mixing efficiency g ¼ 0.2 based on
certain previous work (e.g., Osborn, 1980; Oakey,
1982; PGT-88; Moum et al., 1989; Ruddick et al.,
1997; Lozovatsky et al., 1999; Finnigan et al., 2002;
Lozovatsky and Fernando, 2002; Rippeth and Inall,
2002). Very recently, Oakey and Greenan (2004)
confirmed that, on average, g can be treated as a
constant 0.2, based on a comprehensive set of
microstructure measurements in the coastal waters
of New England. Other studies (e.g., Ivey and
Imberger, 1991; Moum, 1996; Fringer and Street,
2003; Wuest and Lorke, 2003) suggest different g—
ranging between 0.1 and 0.4 (Moum, 1990).
Laboratory experiments of Rohr and Van Atta
(1987) were among the first to show a growing trend
of g with Ri for 0oRio0.25. In addition, according
to the DNS results of Smyth et al. (2001), g can be
time dependent, e.g., growing from 0 to 0.9 as
K–H billows overturn and then decreasing to g0.2
at later stages. Similar observations have been noted
by Peltier and Caulfield (2003) (also see Balmforth
et al., 1998, for the applications of these results).
Recent laboratory results (Strang and Fernando,
2001a) and the GISS mixed model (Canuto et al.,
2001) suggest a continuous growth of g with Ri for
0oRio1 (see Fig. 10), and upon reaching a
maximum g starts to decrease with Ri. The GISS
model employs specific damping functions Sm and
Sh (also called structure functions) for momentum
and temperature (buoyancy) diffusivities K m;s ¼
Sm;s q4 =e; respectively (Fig. 11), which result in a
growing g(Ri) as shown in Fig. 10. These recent
theoretical and laboratory results can be compared
with our microstructure measurements by calculating the diffusivities Kb using an Ri-dependent g.
Fig. 10 shows several examples of g(Ri) obtained
in the laboratory (1), in a stably stratified nocturnal
atmospheric boundary layer (2), and those utilized
in the Mellor–Yamada parameterization Scheme (3)
and in the GISS model (4). The GISS related
samples of g(Ri) and their damping functions
(Fig. 11a) were obtained by digitizing Figs. 5 and 2,
ARTICLE IN PRESS
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
respectively, of Canuto et al. (2001). Continuous
growth of g with Ri for Rio1 is a characteristic
feature of all data presented in Fig. 10. Strang and
Fernando (2001a) measurements exhibit the fastest
399
growth while the Mellor–Yamada function approaches a level of 0.325, corresponding to the flux
Richardson number limit of
Rf ¼ 0:725ðRi þ 0:186 ðRi2 0:316Ri
þ 0:0346Þ1=2 Þ,
ð6Þ
while being equal to 0.245 when Ri-1. The
atmospheric data follow a power fit
g ¼ 0:66Ri2=3
for Rio1
(7)
and the Canuto et al. (2001) dependence can be
approximated by a polynomial function
gC ¼ 0:01 þ Ri 0:53Ri2
(8)
that allows a decrease of g beyond Ri ¼ 1. Using
these approximations for g(Ri) and a 5th-order
polynomial fit to Strang and Fernando (2001a)
data, we have recalculated diffusivities (Fig. 9) by
multiplying the measured values of e=N 2 by g(Ri)
rather than g ¼ 0.2. Fig. 12 shows almost constant
level of K^ b corresponding to Ri-dependent g
calculations for Rio1. By using a polynomial
approximation to the Sh damping function of the
eddy diffusivity:
Fig. 11. (a) The momentum Sm (5) and heat (buoyancy) Sh (6)
damping functions for the diffusivities of Canuto et al. (2001); (b)
the Sh/Sm ratio approximated by Eq. (10a) (7) and Eq. (10b) (8).
S h ðRiÞ ¼ 2ð0:05 0:117Ri þ 0:093Ri2 0:025Ri3 Þ,
(9)
we obtained a good fit to the K^ b ¼ gðRiÞe=N 2
samples shown by a dashed line in Fig. 12. Note
Fig. 12. The normalized buoyancy diffusivities calculated for a constant (0.2) and Ri-dependent mixing efficiency g (details are in the text).
The scalings given by Eq. (4) (a bold line) and Canuto et al. (2001) (a dashed line, Eq. (9)) fit well the two versions of diffusivities. The
confidence limits of the original data shown in Fig. 9 are removed to make the plot more transparent.
ARTICLE IN PRESS
400
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
that the samples at highest and lowest Ri deviate
from the main trend, which could be attributed to
the low statistics at the edge points and insufficient
accuracy of e=N 2 calculations at very low Ri. The
result implies that our measurements of e=N 2
support parameterizations (4) and (5), which are
based on a traditional diffusivity calculation (g is a
constant) as well as parameterization (9), which
accounts for an Ri-dependent g (Eq. 8). The
question, therefore, becomes, which of these approaches leads to a diffusivity that best represents
vertical mixing in numerical boundary layer models?
Evaluation of the results of a series of numerical
experiments that employ each of these parameterizations against a set of high quality field data is
needed to make a definitive judgment. Such a
comprehensive investigation is beyond the framework of the present study, but below we offer a
simple test that sheds light on this issue.
using a standard set of constants (Rodi, 1993),
viz., C1e ¼ 1.44, C2e ¼ 1.92, C3e ¼ 0, Cm ¼ 0.09,
se ¼ 1.3, and sq ¼ 1 (experiments with C3e ¼ 1.4
and 2/3 suggested by Baumert and Peters, 2000,
2004, did not reveal any significant influence of C3e).
The following background vertical profiles were
used: for turbulent kinetic energy q20 ðzÞ ¼
106 m2 =s2 , eddy viscosity K 0 ðzÞ ¼ n ¼ 106 m2 =s,
and the dissipation rate e0 ðzÞ ¼ C m q40 =K 0 .
To incorporate the GISS model approach, we
approximated function Sm (Fig. 11a) as
Sm ¼ 2ð0:05 0:069Ri þ 0:0021Ri2 Þ; Rio0:95
and
Sm ¼ expð2:9RiÞ; Ri40:95
ð10aÞ
and used the inverse turbulent Prandtl number
Pr1
tr ¼ S h =S m in the form
Sh =S m ¼ 1 Ri; Rio0:45
and
6. Mixed-layer modeling
Sh =S m ¼ expð1:3RiÞ; Ri40:45.
The data obtained at St. 946, where MLD was
only 5–6 m, have been used as the initial conditions
for a series of numerical experiments that simulate
the development of an 45 m mixed layer under the
influence of winds of 10–12 m/s. The results of the
calculations are compared with the data from St.
948, where the mixed layer was 48 m deep. St. 946
data were taken under calm winds (3 m/s). The wind
rapidly increased from 3 to 11 m/s over an 12 h
prior to the arrival of the ship at St. 948 (i.e., after
the end of the measurements at St. 947, 30 miles
south of the transect). During the 12 h of sailing
toward St. 948 the wind was almost constant with
u ¼ 1:4 102 m=s, which is close to the transectaveraged u ¼ 1:25 102 m=s.
Assuming, that the horizontal variability of
hydrophysical fields over a distance of 30–60 miles
in the middle of the ocean is insignificant (the 1storder approximation), it is possible to use a onedimensional (1D) boundary layer model to evaluate
the different mixing parameterization schemes discussed in Sections 4 and 5. Considering the
dominance of wind mixing, we may neglect surface
buoyancy forcing. An in-house version of a q2 e
1D model consisting of prognostic momentum
equations for stratified rotational flows as well
as for buoyancy, turbulent kinetic energy q2 and
its dissipation e was used (see, e.g., Rodi, 1993).
The calculations were made with 4 cm vertical
resolution and a 4 s time step. The model was run
Approximations (10b) are shown in Fig. 11b
vis-à-vis the digitized samples of the original Canuto
et al. (2001) ratio S h =Sm .
Fig. 13 shows the initial profile (St. 946) of
density excess over the sea-surface density and the
modeled profiles after 12 h of wind work for the
parameters suggested for Eqs. (2) and (3), and those
based on PP-81 and the GISS models (Fig. 11). The
density excess at St. 948 is also included in the figure
for comparison. The results indicate that the closest
to the observed density profile was achieved by the
GISS mixing model. During 12 h of constant wind
with u ¼ 1:25 102 m=s, an MLD of about 40 m
was developed in agreement with the basic features
of the density profile at St. 948. Other parameterizations used also develop a mixed layer, but of
lesser depth. Almost the same 40 m of the upper
layer was influenced by the diffusivities given by
Eqs. (2) and (3) and PP-81, but below z 20 25 m
incomplete mixing was observed. Approximately
22–24 h of wind work was needed for Eqs. (2) and
(3) to create a density structure similar to that
produced by the GISS model over 12 h.
In the sensitivity runs, the influence of the vertical
step DZ used for the estimates of velocity and density
gradients was verified by changing DZ in the range
4 cm to 8 m (ADCP sampling rate) but keeping all
variables at 4 cm vertical grid and centering the
gradients and the Richardson numbers appropriately
depending on DZ. The modeling density profiles did
not exhibit noticeable differences for DZ ¼ 2, 4, and
ð10bÞ
ARTICLE IN PRESS
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
Fig. 13. Results of numerical calculations using a q2e 1D model
for various values of p and r in Eqs. (2) and (3), Pacanowski and
Philander (1981) formulae (PP-81), and for the GISS model
approximations (10a) and (10b). A constant wind of 11 m/s
worked 12 h to modify the initial density profile measured at St.
946 (before the calculations it was interpolated at Dz ¼ 4 cm).
The modeling profiles of density excess dr over the surface
density are compared with the measured profile of dr at St. 948
(the dotted line; Dz ¼ 2 m).
8 m at the time period between 12 and 24 h after the
onset of calculations. This indicates that the vertical
lengthscale of vertical inhomogeneities of mean Sh,
N2, and Ri exceeds at least 8 m in the weakly
stratified upper layer.
The normalized diffusivity K^ b given by Eqs. (4)
and (5) can be directly utilized for prognostic
calculations of density and velocity while omitting
the equations for q2 and e, the results for which
[Soloviev et al.’s (2001) parameterization with
Ricr ¼ 0.1] are shown in Fig. 14 for 6, 12 and 15 h
of wind work for 11 m/s constant wind. It yields a
deep quasi-homogeneous layer with MLD of about
40 m. The GISS model has slightly more mixing
affinity; its density profile for t ¼ 12 h almost
coincides with that produced by Eq. (5) for
t ¼ 15 h. The Richardson numbers associated with
Eq. (5) continuously grow with depth. The MLD
can be identified by RiE1, but over the next 10 m of
the pycnocline, Ri(z) increases slowly and then
jumps to a high background value. The GISS-based
Ri profile clearly distinguishes the MLD, where Ri
drops to 0.1, but then sharply increases at the
upper boundary of pycnocline. The parameterization based on Eq. (4)—not shown—leads to a
shallower mixed layer due to lower values of Ricr
and Rib.
401
Fig. 14. Modeling density and Richardson number profiles after
6, 12, and 15 h of the wind work using Eq. (5) for Kb
parameterization without q2 and e balance equations. Profiles
of the density excess dr and Ri computed by a complete system of
q2e model equations using GISS parameterization (Eqs. (9) and
(10)) are also shown for comparison.
Finally, we compared the modeled profiles of the
dissipation rate based on the GISS model with that
of Eq. (4), and the results are shown in Fig. 15 (after
12 h of the wind work). The initial dissipation rate
was very small across the entire upper layer (z4zw),
between 2 109 and 8 109 W=kg with the mean
of 4:6 109 W=kg. Turbulence generated by the
constant 11 m/s wind substantially changed the
vertical structure of e(z), producing a dissipation
rate at z ¼ 16 m of more than 106 W/kg that
gradually decreases with depth to 107 W/kg at
zE45 m (disregarding the fine-scale intermittency)
and then sharply drops at z ¼ 50 m to the background level of 5 109 W=kg observed at St. 946
before the wind started. The GISS modeling profile
(e2 in Fig. 15) shows an almost constant level of e
within the mixing layer down to z ¼ 40 m, coinciding with the MLD of the density profile dr2, but
then the turbulence completely vanishes in the
pycnocline. The profile e1 obtained with Eq. (4)
follows the measured dissipation structure e (St.
948) closely down to z ¼ 50 m, exhibiting a gradual
decrease of dissipation in the mixing layer and
penetrating into the pycnocline to sustain developed
ARTICLE IN PRESS
402
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
of the dissipation profile within the mixing layer and
allows the penetration of wind-induced turbulence
into the pycnocline. Note that we have assumed that
the initial density structure at St. 948 is the same as
that at St. 946, and then it was modified by a
constant wind of 11 m/s that blew for a 12 h period.
This is a reasonable, yet not verifiable, assumption. For a more rigorous inter-comparison between
the performances of different mixing parameterizations, a dedicated set of time-dependent CTD,
microstructure and current measurements are
needed.
7. Summary
Fig. 15. The modeling profiles of the dissipation rate (e1 and e2)
and density excess (dr1 and dr2) after 12 h of wind work (11 m/s)
for parameterizations given by Eqs. (2) and (3) and Eqs. (9) and
(10), respectively. The star lines are the measured e(z) profile at
Sts. 946 and 948. The initial density profile at St. 946 is shown in
Fig. 13. The modeling values of e1 and e2 were increased by one
order of magnitude to closely overlay e(948).
active turbulence. Neither parameterization
matches the measured dissipation rate in absolute
values, which could be attributed to the oversimplified nature of the numerical experiments. In
the plot, all modeling profiles were shifted by one
order of magnitude to overlay the measured
dissipation data at St. 948.
At first glance, the modeling test favors the gdependent diffusivities of Strang and Fernando
(2001a) as well as the GISS mixing scheme, if the
sole interest is predicting the mixed-layer depth. The
modified formula (5) of Soloviev et al. (2001),
however, produces a comparable MLD based on
g ¼ 0.2. Nevertheless, considering the availability of
limited data, it is too early to speculate which of the
models are more appropriate, i.e., those relying on a
constant mixing efficiency or traditional Kb(Ri)
scaling [e.g., PP-81 and those given by Eqs. (2)
and (3)]. Specifically, (2)–(3) yields the correct shape
In this paper, the statistics of the turbulent kinetic
energy dissipation rate e and eddy diffusivity Kb in
the upper mixing layer of the North Atlantic were
analyzed and the diffusivity was parameterized
using the wind stress and local gradient Richardson
number Ri.
The wind-stress-induced dissipation ews ¼ u3n =
0:1kLE , where LE ¼ u =f is the Ekman scale,
accounts for more than 90% of the measured mean
dissipation e~ml within the mixing layer. The depth of
the upper weakly stratified layer on the average is
45 m. The probability distribution of e below the
near-surface layer (z416 m) appeared to be lognormal with mean hei ¼ 3:5 107 W=kg and the
median medðeÞ ¼ 1:5 107 W=kg. The Ri distribution function with med(Ri) ¼ 0.1 indicates the
presence of shear instability, and the probability
of Rio0.25 is above 60%. The depth where Ri is
close to 0.25 roughly corresponds to the mixing
layer depth; at the bottom of the mixed-layer Ri is
close to unity.
The eddy diffusivity Kb was quantified as a
function of Ri (Eqs. (2) and (3)). We demonstrated
that the power-law damping function 1/(1+Ri/
Ricr)p for Kb with p ¼ 1 has clear physical meaning.
In stratified flows, it gives K b e=N 2 for RibRicr,
whence the Ozmidov scale LN ðe=N 3 Þ1=2 serves as
the main turbulent lengthscale determining Kb(Ri).
For non-stratified shear flows p ¼ 1 specifies the
shear scale LSh ðe=Sh3 Þ1=2 (Tchen, 1953) as the
governing turbulent lengthscale. In addition, p ¼
2/3 educes the lengthscale LR ¼ ðe=N 2 ShÞ1=2 , which
separates the buoyancy and production subranges
of the spectra of ocean microstructure (Lozovatsky,
1996).
Because turbulence in the upper layer is mainly
driven by the wind stress, the Monin-Obukhov
ARTICLE IN PRESS
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
theory was used to estimate the non-stratified
turbulence in the upper layer by employing the base
diffusivities K sf ðzÞ ¼ kun z for Kb normalization. A
simple parameterization of Kb/Ksf(Ri) by Eq. (4)
successfully collapsed the bin-median estimates of
normalized diffusivity in the range 1:3 102 o
^
Rio5
101 .
Statistical analysis of diffusivities K b ¼ ge=N 2
and the parameterizations of Kb as a function of
Ri relied on a widely used concept of a constant
mixing efficiency, g ¼ 0.2. Conversely, we estimated
Kb using an Ri-dependent mixing efficiency based
on recent laboratory experiments (Strang and
Fernando, 2001a) and the GISS mixing model
(Canuto et al., 2001) that suggest g as growing
functions of Ri for 0oRio1. The diffusivities so
obtained are almost independent on Ri for Rio1, in
contrast to Kb calculated with a constant g that
decreases with increasing Ri. Interestingly, Kb(Ri)
calculated with g ¼ 0.2 as well as the diffusivities of
the GISS model reckoned with variable g agree well
with the same set of data.
To further investigate the differences between
the two approaches, a series of numerical experiments with a standard 1D time-dependent q2 e
model (Rodi, 1993) was conducted to compare
the efficacy of various K(Ri) functions. The development of an 40 m mixed layer during 12 h of
sustained constant wind stress (11 m/s wind) and
zero buoyancy flux was simulated by the GISS
model (with Ri-dependent g), which was in
close agreement with observations. A modified
Soloviev et al. (2001) formula (5) (with constant g ¼ 0.2), however, yielded almost the same
density profile as the GISS model. The scaling
(2)–(3) suggested in this paper produced the correct
general trend of vertical structure for the dissipation
rate within the mixing layer and allowed the
penetration of wind-induced turbulence into the
pycnocline.
The results indicate that from an MLD point of
view our modeling calculations favor the GISS
model, which is also consistent with the laboratory
findings of Strang and Fernando (2001a, b). On the
other hand, a constant g and the suggested
parameterization (2)–(3) are more favorable if the
focus is on vertical structure of microstructure
variables. Further microstructure data and model
evaluations are needed for a more definitive judgment regarding the role of g(Ri) in parameterization
of mixing in sheared, weakly stratified, ocean
boundary layers.
403
Acknowledgements
We wish to thank F. Gomez, L. Montenegro, and
K. Kreyman, who participated in the microstructure
measurements, and the crew of R./V. Akademik
Ioffe. The cruise was organized with financial
support of the Russian Ministry of Science and
Technology. The authors received partial support
from the US Office of Naval Research, grant
N00014-97-1-0140 (I.L. and H.J.S.F.), and NATO
2002 Research Fellowship (I.L.); Spanish Government grant REN2001-2239 and Agencia Catalana
de l’Aigua (E.R.); ANUIES and CICESE, Mexico
(M.F.), and by the Russian Foundation for Basic
Research, grant 02-05-64408 (S.S.).
Appendix A. The MSS profiler
The MSS profiler (Prandke et al., 2000) encapsulates electronics and a set of sensors at the lower end
of a 1 m cylinder. The package is protected by a
guard. The vertical resolution of the airfoil shear
probe (PNS 98) is limited to 2 cm by the sensor
geometry. The lowest in situ noise level of shear
measurements is equivalent to 3 109 W=kg.
The sensitivity of the fast thermistor is 103 C, and
its response time is 7 ms. The precision temperature
sensor (Pt-10) has a sensitivity of 103 C, accuracy
102 C, and a time response of 160 ms. The
corresponding characteristics of the seven-pole
conductivity cell are 103 mS/cm, 102 mS/cm, and
100 ms, respectively. The data are transferred to an
on-board computer via a neutrally buoyant elastic
cable. The profiler operates at a falling speed of
about 0.7 m/s.
Microstructure data are usually contaminated by
ship-induced movements and profiler transients.
Since near-surface data segments cannot be recovered by a denoising procedure, the data at zo16 m
was removed from the analysis. This near-surface
interval can be probed only with uprising (e.g.,
Soloviev et al., 1999) microstructure profilers. Airfoil data in the oceanic upper 15–20 m layer taken
during free fall are usually discarded (e.g., Oakey,
1982; Peters et al., 1988; Hebert et al., 1991; Paka
et al., 1999). Below this heavily contaminated depth
interval, small-scale shear signal was denoised by
removing isolated spikes, bad values, or gaps in the
records identified by an interactive graphic interface. Bad samples were replaced using a cubic spline
interpolation, if the number of bad or missing
points were less than fifty (about 5 cm of the
ARTICLE IN PRESS
404
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
record). A localized, narrow-frequency noise, which
sometimes appear in the signal because of mechanical resonance of the profiler, was eliminated by a
very sharp Lanczos filter (Hamming, 1987) tuned to
a specific frequency band (usually 40 Hz).
After appropriate editing to the small-scale shear
signal was applied, the dissipation rate e was
calculated by fitting 1D wavenumber shear spectra
to the Panchev and Kesich (1969) theoretical
spectrum, following the recommendation of Gregg
et al. (1996). This spectrum contains a little more
power at lower wavenumbers and rolls off slightly
faster at high wavenumbers compared to the wellknown Nasmyth (1970) spectrum. The difference
between the two spectra affects the measurements at
high levels of dissipation, whereupon small-scale
shear at low wavenumbers is not well resolved;
therefore, a correction is needed when the shear
variance is calculated. The experimental spectra
showed a good agreement with the universal spectra
at intermediate wavenumbers, covering the most
important range for our purpose, which includes the
maximum of the dissipation spectrum. The dissipation estimates were obtained at each 1 m vertical
segment, which were then used for Kb calculations
after appropriate averaging.
Appendix B. The Kb(Ri) error estimates
B.1. e, N2, and Kb
The errors of the diffusivity Kb ¼ ge/N2 and
Richardson number estimates depend on the
accuracy of field measurements and computing
algorithms. Assuming that the mixing efficiency g
is a constant (not measured in our case) and e, N,
and vertical shear Sh are statistically independent
variables, it is possible to estimate the relative
(normalized) RMS errors of Kb as
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
(B.1)
dK b ¼ d2e þ d2N 2 ,
where dy ¼ sy =y and s2y is the variance of a random
variable y (Korn and Korn, 1961). The laboratory
calibration test for the shear probe gives a noise
level about 1011 W/kg (Prandke et al., 2000), but
the total computational RMS accuracy of the
dissipation estimates obtained by MSS does not
exceed se ¼ (35) 109 W/kg (Roget et al., 2006).
Given the dissipation rate in the mixing layer was
above 107 W/kg, the normalized RMS error of the
dissipation estimates de ¼ se =e becomes less than
3–5%.
The computational error of N2 is governed by the
RMS error of density measurements sr and the
vertical step Dz of density gradient calculations
pffiffi
2gsr
sN 2
,
(B.2)
¼
2
rDzN 2
N
where sr depends on the RMS noise levels of
temperature sT, conductivity sC, and pressure sp
channels and T, S, and p partial derivatives of
density. According to Prandke et al. (2000), for
MSS sensors, sTE103 K over a bandwidth
0–25 Hz, sCE104 S/m over a bandwidth
0–100 Hz and spE1 GPa. In the weakly stratified
upper QHL (N2 is less than 105 s2), the partial
derivatives are small, salinity spikes due to mismatch of response time between temperature and
conductivity sensors are almost unrecognizable and
therefore the upper limit for sr can be set as
5 104 kg/m3. Thus, the normalized RMS errors
of N2 calculated over 4 m vertical segments using
(B.2) are: dN 2 ¼ sN 2 =N 2 ¼ 0:2 2 in the range
N2 ¼ 105–106/s2. These numbers are relatively
high, showing the possibility of significant uncertainties of individual N2 samples, specifically for
very low N2. To improve the quality of the
estimates, averaging over larger domains or several
profiles should be conducted.
To estimate the error dK b of diffusivity from
(B.1), we used de ¼ 0:05, which is much less than
dN 2 , and therefore dK b is about the same as dN 2 ,
ranging between 0.2–0.25 and 2.
B.2. Shear and Ri
The uncertainty of shear measurements is determined by the nominal RMS error of a single bin,
which is 30 cm/s for 8 m cell of 75 kHz ADCP, and
the number of bins used to obtain an averaging
estimate of velocity at a particular depth. Because
our measurements were taken only at drift stations
(typical drift 1–1.5 knots), the errors related to
ship movement were small (Firing, 1988). Modern
GPS navigation allowed quite accurate estimates for
the absolute velocity profile to be obtained (Lien et
al., 1994). For the estimates of vertical shear, the
accuracy of absolute currents, however, is not
important. Individual profiles of horizontal current
components uðzÞ and vðzÞ were obtained for 18 min
averaging intervals, which reduced the nominal
ARTICLE IN PRESS
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
velocity error by a factor of 40. Then, the individual
profiles were averaged over the measurement time at
each station (4–7 profiles during about 1.5–2 h) and
~
such averaged profiles uðzÞ
and v~ðzÞ were used to
calculate the shear magnitude Sh ¼ jdV =dzj ¼
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
~
ðdu=dzÞ
þ ðd~v=dzÞ2 . The RMS error of jV j ¼
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u~ 2 þ v~2 is about 0.5 cm/s, producing an RMS
error of Sho0.001/s. This is consistent, e.g., with
the corresponding estimates reported by Peters et al.
(1988). Typical magnitude of the shear in QHL was
in the range 5 103 2 102 s1 (see Fig. 7, for
example). Therefore, the RMS normalized error of
the shear estimates is in the range 5–20% and for
the squared shear it goes up to 10–30%.
Note that a composite spectrum of vertical shear
(Gargett et al., 1981) contains a turbulence subrange
bounded at low wavenumbers by the Ozmidov scale
LN, which is about 1 m in the ocean thermocline.
Therein the characteristic scales of local background shear and stratification are on the order of
LN. In QHL, however, the Ozmidov scale may grow
to several 10 m (see, e.g., Fig. 7), whence other
lengthscales become dominant. For example, the
largest scales of 3D turbulent eddies in QHL are
determined by the external turbulent scale L0, which
is of the order of ð0:1 0:2Þhe (Monin and Yaglom,
1975); most recently Canuto et al. (2002) identified
L0 as 0:17hD . Because in our measurements the
mean thickness of QHL hhD i hhe i ¼ 45 m, the
characteristic L08 m. As such, the 8 m resolution
ADCP shear is appropriate to evaluate the Richardson number pertinent to the QHL small-scale
turbulence. Some support for this conclusion is
provided by our modeling efforts (Section 6), in
which we tested the influence of the vertical
separation DZ on the model outcome by calculating
velocity and density gradients and thus Richardson
numbers with DZ ¼ 2, 4, and 8 m. The modeling
output was almost insensitive to DZ, indicating that
the shear and density structures in the upper weakly
stratified layer did not have substantial variations
on scales o8 m. Therefore, the calculation of Ri
with an 8 m vertical resolution seems appropriate.
The relative RMS error for the squared shear dSh2
is considerably lower in most cases than dN 2 .
Therefore, the relative error of Ri is mainly
determined by dN 2 rather than dSh2 , although a
combination of relatively high N2 and Sh2 can be
encountered near the base of the upper mixing layer.
The estimate of dRi could be as high as 3, but the
probability of such a high uncertainty level is low
given that small shear in the upper turbulent layer
405
was mostly observed near the lower boundary with
an underlying pycnocline where N2 starts to
increase. Therefore, higher dSh2 is compensated by
lower dN 2 and dRi is expected to remain at an
acceptable intermediate range. In order to increase
the confidence level on the dependence between Ri
and Kb, we performed equal-probability bin averaging, which reduced the uncertainty of the
averaged samples shown in Fig. 9 by a factor of 3.
References
Balmforth, N.J., Llewellyn Smith, S.G., Young, W.R., 1998.
Dynamics of interface and layers in a stratified turbulent fluid.
Journal of Fluid Mechanics 355, 329–358.
Baumert, H., Peters, H., 2000. Second-moment closures and
length scales for weakly stratified turbulent shear flows.
Journal of Geophysical Research 105, 6453–6468.
Baumert, H., Peters, H., 2004. Turbulence closure, steady sate
and collapse into waves. Journal of Physical Oceanography
34, 505–509.
Brainerd, K.E., Gregg, M.C., 1995. Surface mixed and mixing
layer depths. Deep-Sea Research I 42, 1521–1544.
Canuto, V.M., Howard, A., Cheng, Y., Dubovikov, M.S., 2001.
Ocean turbulence. Part I: one-point closure model—momentum and heat vertical diffusivities. Journal of Physical
Oceanography 31, 1413–1426.
Canuto, V.M., Howard, A., Cheng, Y., Dubovikov, M.S., 2002.
Ocean turbulence. Part II: vertical diffusivities of momentum,
heat, salt, mass, and passive scalars. Journal of Physical
Oceanography 32, 240–264.
Chereskin, T.K., Firing, E., Gast, J.A., 1989. On identifying and
screening filter skew and noise bias in acoustic Doppler
current profiler measurements. Journal of Atmospheric and
Oceanic Technology 6, 1040–1054.
Finnigan, T.D., Luther, D.S., Lukas, R., 2002. Observations of
enhanced diapycnal mixing ner the Hawaiian Ridge. Journal
of Physical Oceanography 32, 2988–3002.
Firing, E., 1988. Report from the WOCE/NOAA Workshop on
ADCP measurements, Austin, TX, March 1–2, 1988. US
WOCE Planning Report No. 13, 97pp, US Planning Office
for WOCE, College Station, TX.
Fringer, O.B., Street, R.L., 2003. The dynamics of breaking
progressive interfacial waves. Journal of Fluid Mechanics 494,
319–353.
Gargett, A.E., Hendricks, P.J., Sanford, T.B., Osborn, T.R.,
Williams, A.J., 1981. A composite spectrum of vertical shear
in the upper ocean. Journal of Physical Oceanography 11,
1258–1271.
Gargett, A.E., Osborn, T.R., Nasmyth, P.W., 1984. Local
isotropy and the decay of turbulence. Journal of Fluid
Mechanics 144, 231–280.
Gibson, C.H., 1980. Fossil temperature, salinity and vorticity
turbulence in the ocean. In: Nihoul, J.C.J. (Ed.), Marine
Turbulence. Elsevier, Amsterdam, pp. 221–257.
Gregg, M.C., Winkel, D.P., Sanford, T.B., Peters, H., 1996.
Turbulence produced by internal waves in the oceanic
thermocline at mid and low latitudes. Dynamics of Atmospheres and Oceans 24, 1–14.
ARTICLE IN PRESS
406
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
Gurvich, A.C., Yaglom, A.M., 1967. Breakdown of eddies and
probability distributions for small-scale turbulence. Physics of
Fluids 10, S59.
Hamming, R.W., 1987. Numerical Methods for Scientists
and Engineers, second ed. Dover Publications, New York
(721pp).
Hebert, D., Moum, J.N., Caldwell, D.R., 1991. Does ocean
turbulence peak at the equator?: revisited. Journal of Physical
Oceanography 21, 1690–1698.
Henderson-Sellers, B., 1982. A simple formula for vertical eddy
diffusion coefficients under conditions of non-neutral stability. Journal of Geophysical Research 87, 5860–5864.
Ivey, G.N., Imberger, J.J., 1991. On the nature of turbulence in a
stratified fluid: I, the energetics of mixing. Journal of Physical
Oceanography 21, 650–658.
Jacobitz, F.G., Sarkar, S., Van Atta, C.W., 1997. Direct
numerical simulation of the turbulence evolution in a
uniformly sheared and stably stratified flow. Journal of Fluid
Mechanics 342, 231–261.
Kara, A.B., Rochford, P.A., Hurlburt, H.E., 2000. An optimal
definition for ocean mixed layer depth. Journal of Geophysical Research 105, 16803–16821.
Korn, G.A., Korn, T.M., 1961. Mathematical Handbook for
Scientists and Engineers, McGraw-Hill, New York (720pp).
Lee, S.M., Giori, W., Princevac, M., Fernando, H.J.S., 2005. A
new turbulent parameterization for the nocturnal PBL over
complex terrain. Boundary Layer Meteorology, in press.
Lien, R.-C., McPhaden, M.J., Hebert, D., 1994. Intercomparison
of ADCP measurements at 01, 1401W. Journal of Atmospheric and Oceanic Technology 11, 1334–1349.
Linden, P.F., 1979. Mixing in stratified fluids. Geophysical and
Astrophysical Fluid Dynamics 13, 3–23.
Lombardo, C.P., Gregg, M.C., 1989. Similarity scaling of viscous
and thermal dissipation in a convecting surface boundary
layer. Journal of Geophysical Research 94, 6273–6284.
Lozovatsky, I.D., 1996. Turbulence decay in stratified and
homogeneous marine layers. Dynamics of Atmospheres and
Oceans 24, 15–25.
Lozovatsky, I.D., Fernando, H.J.S., 2002. Mixing on a shallow
shelf of the Black Sea. Journal of Physical Oceanography 32,
945–956.
Lozovatsky, I.D., Ksenofontov, A.S., Erofeev, A.Yu., Gibson,
C.H., 1993. Modeling of evolution of vertical structure in the
upper ocean by atmospheric forcing and intermittent turbulence in the pycnocline. Journal of Marine Systems 4,
263–273.
Lozovatsky, I.D., Dillon, T.M., Erofeev, A.Yu., Nabatov, V.N.,
1999. Variations of thermohaline structure and turbulent
mixing on the Black Sea shelf at the beginning of autumn
cooling. Journal of Marine Systems 21, 255–282.
Lozovatsky, I.D., Ksenofontov, A.S., Fernando, H.J.S., 2000.
The formation of step-like structure in near surface and nearbottom pycnoclines. In: Lawrence, G., Pieters, R., Yonemitsu, N. (Eds.), Stratified Flows II. UofBC, Vancouver,
Canada, pp. 1215–1220.
Lozovatsky, I.D., Roget, E., Figueroa, M., Fernando, H.J.S.,
Shapovalov, S., 2005. Observations and scaling of the upper
mixed layer in the North Atlantic. Journal of Geophysical
Research 110, C05013.
Mellor, G.L., Yamada, T., 1982. Development of a turbulence
closure model for geophysical fluid problems. Reviews of
Geophysics and Space Physics 20, 851–875.
Millard Jr., R.C., Lake, B.J., Brown, N.L., Tool, J.M., Schaaf,
D., Yang, K., Yu, H., Zhou, L., 1990. US/PRC CTD
Intercalibration Report 1986–1990. WHOI Technical Report
90–53.
Monin, A.S., Obukhov, A.M., 1954. Basic turbulent mixing laws
in the atmospheric surface layer. Trudy Geofizicheskogo
Instituta AN SSSR 24, 163–187.
Monin, A.S., Yaglom, A.M., 1975. Statistical Fluid Mechanics:
Mechanics of Turbulence, vol. 2. MIT Press, Cambridge, MA
(874pp).
Monti, P., Fernando, H.J.S., Chan, W.C., Princevac, M.,
Kowalewski, T.A., Pardyjak, E., 2002. Observations of flow
and turbulence in the nocturnal boundary layer over a slope.
Journal of Atmospheric Science 59, 2513–2534.
Moum, J.N., 1990. The quest for Kr—preliminary results from
direct measurements of turbulent fluxes in the ocean. Journal
of Physical Oceanography 20, 1980–1984.
Moum, J.N., 1996. Efficiency of mixing in the main thermocline.
Journal of Geophysical Research 101, 12057–12069.
Moum, J.N., 1998. Quantifying vertical fluxes from turbulence in
the ocean. Oceanography 10, 111–115.
Moum, J.N., Osborn, T.R., Paulson, C.A., 1989. Mixing in the
equatorial surface layer. Journal of Geophysical Research 94,
2005–2021.
Munk, W.H., Anderson, E.R., 1948. Notes on the theory of the
thermocline. Journal of Marine Research 3, 276–295.
Nasmyth, P.W., 1970. Oceanic Turbulence. Ph.D. Dissertation,
University of British Columbia, Vancouver.
Oakey, N.S., 1982. Determination of the rate of dissipation of
turbulent energy from simultaneous temperature and velocity
shear microstructure measurements. Journal of Physical
Oceanography 12, 256–271.
Oakey, N.S., 1985. Statistics of mixing parameters in the upper
ocean during JASIN Phase II. Journal of Physical Oceanography 15, 1662–1675.
Oakey, N.S., Greenan, B.J.W., 2004. Mixing in a coastal
environment: 2. A view from microstructure measurements.
Journal of Geophysical Research 109, C10014.
Osborn, T.R., 1980. Estimates of the local rate of vertical
diffusion from dissipation measurements. Journal of Physical
Oceanography 10, 83–89.
Pacanowski, R.C., Philander, S.G.H., 1981. Parameterization of
vertical mixing in numerical models of tropical oceans.
Journal of Physical Oceanography 11, 1443–1451.
Paka, V.T., Nabatov, V.N., Lozovatsky, I.D., Dillon, T.M.,
1999. Ocean microstructure measurements by BAKLAN and
GRIF. Journal of Atmospheric and Oceanic Technology 16,
1519–1532.
Panchev, S., Kesich, D., 1969. Energy spectrum of isotropic
turbulence at large wavenumbers. Comptes Rendus de
l’Académie Bulgare des Sciences 22, 627–630.
Pardyjak, E.R., Monti, P., Fernando, H.J.S., 2002. Flux
Richardson number measurements in stable atmospheric
shear flows. Journal of Fluid Mechanics 459, 307–316.
Pelegri, J.L., Csanady, G.T., 1994. Dyapicnal mixing in western
boundary currents. Journal of Geophysical Research 99,
18275–18304.
Peltier, W.R., Caulfield, C.P., 2003. Mixing efficiency in stratified
shear flows. Annual Review of Fluid Mechanics 35, 135–167.
Peters, H., Gregg, M.C., Tool, J.M., 1988. On the parameterization of equatorial turbulence. Journal of Geophysical
Research 93, 1199–1218.
ARTICLE IN PRESS
I.D. Lozovatsky et al. / Deep-Sea Research I 53 (2006) 387–407
Prandke, H., Holtsch, K., Stips, A., 2000. MITEC technology
development: the microstructure/turbulence measuring system MSS. Technical Report EUR 19733 EN, European.
Rippeth, T.P., Inall, M.E., 2002. Observations of the internal tide
and associated mixing across the Malin Shelf. Journal of
Geophysical Research 107, 3028.
Rodi, W., 1993. Turbulence Models and their Application in
Hydraulics: a State-of-the-Art Review, third ed. International
Association of Hydraulic Research, Balkema.
Roget, E., Lozovatsky, I.D., Sánchez, X., Figueroa, M., 2006.
Small-scale shear and temperature measurements in natural
systems: methodology and applications. Progress in Oceanography, in press.
Rohr, J., Van Atta, C., 1987. Mixing efficiencies in stably
stratified growing turbulence. Journal of Geophysical Research 92, 5481–5488.
Ruddick, B., Walsh, D., Oakey, N., 1997. Variations in apparent
mixing efficiency in the North Atlantic Central Water.
Journal of Physical Oceanography 27, 2589–2605.
Smyth, W.D., Moum, J.N., Caldwell, D.R., 2001. The efficiency
of mixing in turbulent patches: inferences from direct
simulations and microstructure observations. Journal of
Physical Oceanography 31, 1969–1992.
Soloviev, A., Lukas, R., Hacker, P., Baker, M., Schoeberlein, H.,
Arjannikov, A., 1999. A near-surface microstructure sensor
system used during TOGA COARE. Part II: turbulence
measurements. Journal of Atmospheric and Oceanic Technology 16, 1598–1618.
407
Soloviev, A., Lukas, R., Hacker, P., 2001. An approach to
parameterization of the oceanic turbulent boundary layer in
the western Pacific warm pool. Journal of Geophysical
Research 106, 4421–4435.
Stillinger, D.C., Helland, K.N., Van Atta, C.W., 1983. Experiments on the transition of homogeneous turbulence to
internal waves in a stratified fluid. Journal of Fluid Mechanics
131, 91–122.
Strang, E.J., Fernando, H.J.S., 2001a. Vertical mixing and
transports through a stratified shear layer. Journal of Physical
Oceanography 31, 2026–2048.
Strang, E.J., Fernando, H.J.S., 2001b. Entrainment and mixing in
stratified shear flows. Journal of Fluid Mechanics 428, 349–386.
Tchen, C.M., 1953. On the spectrum of energy in turbulent shear
flow. Journal of Research of the National Bureau of
Standards 50, 51–62.
Tereschenkov, V.P., Shapovalov, S.M., Dobrolyubov, S.A., Morozov, E.G., 2002. Cruise 9 of R/V Akademik Ioffe. Oceanology
42, 298–301 (translated from Okeanologiya 42, 315–318).
Toorman, E.A., 2000. Stratification in fine-grained sedimentladen turbulent flows. In: Lawrence, G., Pieters, R.,
Yonemitsu, N. (Eds.), Stratified Flows II. UofBC, Vancouver,
Canada, pp. 945–950.
Ueda, H., Mitsumoto, S., Komori, S., 1981. Buoyancy effects on the
turbulent transport processes in the lower atmosphere. Quarterly
Journal of the Royal Meteorological Society 107, 561–578.
Wuest, A., Lorke, A., 2003. Small-scale hydrodynamics in lakes.
Annual Review of Fluid Mechanics 35, 373–412.