0% found this document useful (0 votes)
8 views10 pages

On The Flux Richardson Number in Stably Stratified Turbulence

Uploaded by

Steve Okumu
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
Download as pdf or txt
0% found this document useful (0 votes)
8 views10 pages

On The Flux Richardson Number in Stably Stratified Turbulence

Uploaded by

Steve Okumu
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
Download as pdf or txt
Download as pdf or txt
You are on page 1/ 10

J. Fluid Mech. (2016), vol. 798, R1, doi:10.1017/jfm.2016.

340

On the flux Richardson number in stably


stratified turbulence

Subhas K. Venayagamoorthy1,2, † and Jeffrey R. Koseff2


1 Department of Civil and Environmental Engineering, Colorado State University, Fort Collins,
CO 80523-1372, USA
2 The Bob and Norma Street Environmental Fluid Mechanics Laboratory,
Department of Civil and Environmental Engineering, Stanford University, Stanford,
CA 94305-4020, USA

(Received 5 April 2016; revised 19 April 2016; accepted 12 May 2016;


first published online 8 June 2016)

The flux Richardson number Rf (often referred to as the mixing efficiency) is a


widely used parameter in stably stratified turbulence which is intended to provide a
measure of the amount of turbulent kinetic energy k that is irreversibly converted to
background potential energy (which is by definition the minimum potential energy
that a stratified fluid can attain that is not available for conversion back to kinetic
energy) due to turbulent mixing. The flux Richardson number is traditionally defined
as the ratio of the buoyancy flux B to the production rate of turbulent kinetic energy
P. An alternative generalized definition for Rf was proposed by Ivey & Imberger
(J. Phys. Oceanogr., vol. 21, 1991, pp. 650–658), where the non-local transport
terms as well as unsteady contributions were included as additional sources to
the production rate of k. While this definition precludes the need to assume that
https://doi.org/10.1017/jfm.2016.340 Published online by Cambridge University Press

turbulence is statistically stationary, it does not properly account for countergradient


fluxes that are typically present in more strongly stratified flows. Hence, a third
definition that more rigorously accounts for only the irreversible conversions of energy
has been defined, where only the irreversible fluxes of buoyancy and production
0
(i.e. the dissipation rates of k and turbulent potential energy (EPE )) are used. For
stationary homogeneous shear flows, all of the three definitions produce identical
results. However, because stationary and/or homogeneous flows are typically not
found in realistic geophysical situations, clarification of the differences/similarities
between these various definitions of Rf is imperative. This is especially true given
the critical role Rf plays in inferring turbulent momentum and heat fluxes using
indirect methods, as is commonly done in oceanography, and for turbulence closure
parameterizations. To this end, a careful analysis of two existing direct numerical
simulation (DNS) datasets of stably stratified homogeneous shear and channel flows
was undertaken in the present study to compare and contrast these various definitions.
We find that all three definitions are approximately equivalent when the gradient

† Email address for correspondence: vskaran@colostate.edu


c Cambridge University Press 2016 798 R1-1
S. K. Venayagamoorthy and J. R. Koseff
Richardson number Rig 6 1/4. Here, Rig = N 2 /S2 , where N is the buoyancy frequency
and S is the mean shear rate, provides a measure of the stability of the flow. However,
when Rig > 1/4, significant differences are noticeable between the various definitions.
In addition, the irreversible formulation of Rf based on the dissipation rates of k and
0
EPE is the only definition that is free from oscillations at higher gradient Richardson
numbers. Both the traditional definition and the generalized definition of Rf exhibit
significant oscillations due to the persistence of linear internal wave motions and
0
countergradient fluxes that result in reversible exchanges between k and EPE . Finally,
we present a simple parameterization for the irreversible flux Richardson number R∗f
based on Rig that produces excellent agreement with the DNS results for R∗f .

Key words: geophysical and geological flows, stratified turbulence, turbulent mixing

1. Introduction
The quantification of irreversible vertical turbulent mixing of momentum and
density in stably stratified geophysical flows in both the Earth’s hydrosphere and
its atmosphere remains an important ongoing challenge. This is not surprising given
the complexity introduced into most geophysical flows by factors such as density
stratification, complex topography and the different physical phenomena associated
with such flows. However, accurate prediction of small-scale irreversible turbulent
mixing is crucial for many practical applications ranging from air quality prediction
in the atmospheric boundary layer to the prediction of heat fluxes and circulation in
oceanic flows. The traditional approach often used for quantifying vertical mixing
in turbulent flows involves specification of the turbulent (eddy) viscosity Km and
diffusivity Kρ which are based on the gradient-transport hypothesis (Pope 2000). For
example, in a unidirectional shear flow, these are given as

u0 w0 

Km = − ,
dU/dz 

(1.1)
ρ 0 w0 
Kρ = − ,

dρ/dz

https://doi.org/10.1017/jfm.2016.340 Published online by Cambridge University Press

where u0 w0 is the Reynolds stress (turbulent momentum flux), ρ 0 w0 is the turbulent


density flux, U is the mean streamwise velocity and ρ is the fluid mean density. Here,
an overbar ( ) implies averaging (spatial or temporal), and the ratio Prt = Km /Kρ
is commonly referred to as the turbulent Prandtl (or Schmidt) number (Launder
& Spalding 1972). Direct estimates of Km and Kρ can be obtained as long as
the turbulent fluxes and the mean gradients can be computed from measurements
(Lozovatsky & Fernando 2013). However, while it is, in principle, possible to obtain
direct estimates of fluxes if certain technical challenges can be resolved (e.g. spatial
resolution issues, especially in water bodies such as the oceans), the coexistence of
internal wave motions and turbulence in stably stratified flows further complicates
such measurements. Separation of flux contributions from linear internal waves (which
have negligible contributions to irreversible mixing) and turbulence still remains a
major difficulty to resolve in flux measurements. Hence, a number of indirect methods
have been used by oceanographers to infer heat and momentum fluxes, such as those
due to Osborn & Cox (1972) and Osborn (1980).
798 R1-2
Flux Richardson number
Estimates of turbulent viscosities and diffusivities from indirect methods essentially
assume that the turbulent flow is statistically stationary (i.e. the turbulence statistics are
invariant in time) and homogeneous (the statistics are invariant under translation). For
example, the popular Osborn (1980) model uses the above assumptions to simplify
the turbulent kinetic energy (k) equation to obtain an estimate for the vertical eddy
diffusivity of density as

Rf 
 
Kρ = , (1.2)
1 − Rf N 2

where  is the dissipation rate of k, N = −(g/ρ0 )∂ρ/∂z is the buoyancy frequency
of the background (stable) density field, where in g is the acceleration due to
gravity, and Rf , the flux Richardson number which provides a measure of the mixing
efficiency, is defined as
B
Rf = , (1.3)
P
where B = g/ρ0 (ρ 0 w0 ) is the buoyancy flux and P = −u0 w0 (dU/dz) is the rate of
production of k. The term Γ = Rf /(1 − Rf ) is also sometimes loosely referred to as the
mixing efficiency in the literature. However, in this paper, we shall refer exclusively to
Rf (together with the two other alternative formulations that will be described below)
as the mixing efficiency. Osborn (1980) recommended a canonical value of Rf ≈ 0.17
(i.e. Γ ≈ 0.2) on the basis of some controlled laboratory experiments by Britter (1974)
and theoretical predictions by Ellison (1957). Since then, numerous studies have been
carried out to ascertain the variability of Rf on the strength of the stratification, which
is usually quantified in terms of the gradient Richardson number Rig = N 2 /S2 , where
S = dU/dz is the mean shear rate (Gregg 1987; Rohr et al. 1988; Fernando 1991;
Ivey & Imberger 1991; Itsweire et al. 1993; Peltier & Caulfield 2003; Ivey, Winters &
Koseff 2008; Shih et al. 2005, among others). However, there is a lack of a universal
parameterization of the mixing efficiency due to both ambiguities inherent in single
parameter based formulations as well as the general complexity of stably stratified
flows (Mater & Venayagamoorthy 2014).
The inherent assumptions of stationarity and homogeneity are idealizations that
are not always applicable from a practical standpoint, be it in observational studies
https://doi.org/10.1017/jfm.2016.340 Published online by Cambridge University Press

or direct numerical simulations (DNS). In an attempt to remedy this issue, Ivey


& Imberger (1991) proposed an alternative generalized definition for Rf (hereafter
denoted by RIIf ) given by
B B
RIIf = = , (1.4)
m B+
where the denominator m now, in addition to P, includes the inhomogeneous transport
terms in the turbulent kinetic equation (i.e. viscous diffusion of k, the divergence of
pressure flux, and the divergence of the turbulent convection of k) and the unsteadiness
term. It should be noted that for stationary homogeneous shear flows, RIIf is identical
to Rf given by (1.3). While this alternative definition precludes the need to assume
that the turbulence is stationary and/or homogeneous, it also suffers from the effects
of countergradient fluxes that are common in more strongly stratified flows. Thus, a
third definition of Rf (denoted commonly as R∗f ) that more rigorously only accounts
for the irreversible conversions of energy locally has recently been defined (Peltier
& Caulfield 2003; Venayagamoorthy & Stretch 2010), where only the irreversible
components of the buoyancy flux (i.e. the dissipation rate PE of the turbulent
798 R1-3
S. K. Venayagamoorthy and J. R. Koseff
0
(available) potential energy EPE ) and of the production term (i.e. the total dissipation
rate,  + PE ) are used, as follows:
PE
R∗f = . (1.5)
 + PE

Here, PE = N 2 ρ (dρ/dz)−2 , in which ρ = κ(∇ρ 0 )2 is the dissipation rate of density


(scalar) variance, with κ defined as the molecular diffusivity of density. Further
details and a conceptual explanation for the basis of this definition can be found
in Venayagamoorthy & Stretch (2010). Again, it is worth noting that for stationary
homogeneous flows, R∗f = RIIf = Rf . However, as discussed earlier, such conditions
are rarely applicable in practice. Given the fact that the data used to develop
parameterizations of the flux Richardson number are from stably stratified flows that
are evolving in time and/or spatially inhomogeneous, it is imperative to investigate the
differences and similarities between these three commonly used definitions of the flux
Richardson number. Such an analysis is likely to provide insights that will not only
be helpful for improved estimates of irreversible fluxes using indirect measurement
techniques, but will also aid the development of more robust parameterizations for
quantifying irreversible mixing in numerical simulations of geophysical flows. We
note that Karimpour & Venayagamoorthy (2015) carried out a detailed analysis
of turbulent mixing in stably stratified channel flow where they also performed a
comparison of two of the three definitions for the flux Richardson number, namely
Rf and R∗f . However, to the best of our knowledge, an in-depth comparative study
involving all three definitions of the flux Richardson number using different flow
configurations has not been carried out to date. This simple and important goal is the
focus of this paper.
In what follows, in § 2, we present a brief description of the datasets that we use
in our study to facilitate such a comparison. The results of our study are presented
in § 3 and conclusions are given in § 4.

2. Data sources
Our objective is to explore the behaviour of the three different definitions of
the flux Richardson number outlined in § 1 using available data that have all the
https://doi.org/10.1017/jfm.2016.340 Published online by Cambridge University Press

necessary variables for computing the flux Richardson number as a function of the
gradient Richardson number Rig . Accordingly, consideration is given to datasets that
are obtained for evolving homogeneous and inhomogeneous flows that include mean
shear and density stratification. We primarily use the dataset from the DNS study of
homogeneous shear flows by Shih et al. (2005, hereafter SKIF). These simulations are
for temporally developing homogeneous turbulence subjected to different values of the
uniform mean shear rate and uniform stable stratification. The gradient Richardson
numbers used in the simulations were 0.05 6 Rig 6 1. For low Rig , the turbulent
kinetic energy k grows in time, while at high Ri (>0.25), k decays. An approximately
stationary state was achieved for Ri ' 0.17 for the simulations cases considered here.
All statistics were obtained by volume averaging over the computational domain.
The data presented are for non-dimensional shear time St > 4 to the end of the
simulation time, in order to filter out the initial transients during the development
phase of the turbulence encountered in the initialization of the simulations. Further
details on the simulations can be found in SKIF. We also use the stably stratified
turbulent channel flow DNS dataset of García-Villalba & del Álamo (2011, hereafter
GVA). For the purpose of this study, we use data from simulations performed at a
798 R1-4
Flux Richardson number
1.0
0.8
0.6
0.4
0.2
0
–0.2
–0.4
–0.6
–0.8
–1.0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

F IGURE 1. Comparison of the flux Richardson numbers Rf , RIIf and R∗f as functions of
the gradient Richardson number Rig , computed from the DNS data of Shih et al. (2005).

friction Reynolds number of Reτ = uτ δ/ν = 550, with an initial stratification given
by a friction Richardson number of Riτ = |1ρ|gδ/ρ0 u2τ = 60. Here, uτ is the friction
velocity, δ is half of the channel depth, ν is the kinematic (molecular) viscosity and
|1ρ| is the initial density difference between the bottom of the channel (z = 0) and
the free stream (z = δ).

3. Results
3.1. Comparisons of the three different flux Richardson numbers
Figure 1 shows the flux Richardson numbers Rf , RIIf and R∗f as defined by (1.3), (1.4)
and (1.5) respectively as functions of the gradient Richardson number Rig using the
DNS data of SKIF. There are two main observations worth noting. First, for low
Rig (up to Rig ' 0.25), it can be seen that all three definitions are approximately
equivalent (albeit with some differences which on average are within ±25 %) in this
https://doi.org/10.1017/jfm.2016.340 Published online by Cambridge University Press

shear dominated flow regime. Conceptually, this implies that both the Reynolds stress
and the buoyancy flux are downgradient. Previous studies by Rogers, Mansour &
Reynolds (1989) and Holt, Koseff & Ferziger (1992) postulated that the fluxes in the
shear dominated regime are produced by pumping of fluid through coherent hairpin
shaped eddies. The data suggest that the flux Richardson number (regardless of which
definition is used) increases with Rig in a quasi-linear manner for small Rig (. 0.1)
and continues to increase with decreasing slope for 0.1 . Rig 6 0.25. The favourable
comparison between the three different definitions is encouraging in the sense that
it highlights the fact that both buoyancy and momentum fluxes are dominated by
turbulent processes in this regime. The good agreement suggests that any of the three
definitions could be used for inferring the irreversible mixing in the shear dominated
regime in stably stratified flow with a reasonable degree of accuracy. As a side
note, it is worth mentioning that it is common practice to consider Rig = 1/4 as
the threshold for the onset of shear instabilities in stably stratified flows, and hence
this value is typically termed as the critical Richardson number Rigc (Howard 1961;
Miles 1961). It is interesting to note that the transition in the behaviour of the three
definitions is conceptually consistent with the notion of a critical threshold in Rig .
798 R1-5
S. K. Venayagamoorthy and J. R. Koseff
1.0
0.1
0.8
0.2
0.6
0.3
0.4

0.2 0.4

0 0.5

–0.2 0.6

–0.4 0.7

–0.6 0.8

–0.8 0.9

–1.0 1.0
0 0.5 1.0 1.5 2.0 2.5 3.0

F IGURE 2. Comparison of the flux Richardson numbers Rf , RIIf and R∗f as functions of
the buoyancy Reynolds number Reb , computed from the DNS data of Shih et al. (2005).
The colour bar shows the corresponding gradient Richardson number Rig .

However, it is important to note that this does not imply that a physical transition in
the flow regime has to occur strictly at Rig = 1/4 and thus should only be considered
as a conceptual delineation of flow regimes.
Second, it can be seen that both Rf and RIIf exhibit significant variability at higher
Rig (>0.25), while R∗f shows almost negligible variation in comparison to the former
two definitions. It should be noted that for a given Rig , the variability in Rf indicates
oscillations in time of both the buoyancy flux and the momentum flux, while for
RIIf , the variability signifies oscillations of the buoyancy flux. Physically, at high
Rig (i.e. the so-called buoyancy dominated flow regime), the production P of k
becomes small while the buoyancy flux B is increasingly dominated by adiabatic
displacements from linear internal waves. Hence, the traditional definition of Rf
becomes less meaningful in such buoyancy dominated flow regimes. Similarly, RIIf
https://doi.org/10.1017/jfm.2016.340 Published online by Cambridge University Press

suffers from the significant oscillations at high Rig . However, it is worth noting
that unlike the trend in Rf , on average, RIIf appears to decrease for Rig > 0.25 but
becomes quickly less meaningful at higher Rig values (i.e. attains negative values)
when significant countergradient buoyancy fluxes become prevalent. On the other
hand, the irreversible flux Richardson R∗f does not suffer from such issues since, by
definition, it excludes the effects of reversible contributions. Furthermore, R∗f appears
to asymptote to an approximate constant for high Rig . The approximate constancy of
the overall mixing efficiency computed from time-integrated energy budgets is also
evident at large Richardson numbers in previous studies (Stretch et al. 2010). These
results underscore the importance of separating out the reversible contributions to both
the momentum and scalar fluxes in such time evolving (i.e. locally non-equilibrium)
flows, especially in the strongly stratified buoyancy dominated flow regime.
Given the widespread use of other parameters such as the buoyancy Reynolds
number Reb = /(νN 2 ), where ν is the kinematic viscosity of the fluid (e.g. Shih
et al. 2005), for parameterizing mixing in stratified turbulence, it is instructive to
explore the variation of Rf , RIIf and R∗f with such a parameter. Figure 2 shows the
dependence of Rf , RIIf and R∗f with Reb . It can be seen that all three definitions
798 R1-6
Flux Richardson number
1.0

0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

F IGURE 3. Comparison of the flux Richardson numbers Rf , RIIf and R∗f as functions of
the gradient Richardson number Rig , computed from the DNS data of García-Villalba &
del Álamo (2011).

seem to be in approximate agreement for Reb & 30. This is also consistent with the
low-Rig regime (.0.25), as can be seen in the colour bar insert in figure 2 and the
dependence on Rig shown in figure 1. Furthermore, at lower Reb values (<30, and
correspondingly higher Rig values), the agreement deteriorates similarly to that seen in
figure 1 with Rig . This figure also underscores the fact that turbulence is suppressed
at higher Rig as buoyancy effects become dominant (at least in these simulations).
Figure 3 shows a comparison of Rf , RIIf and R∗f using the DNS data of GVA for
a fully developed stably stratified turbulent channel flow. Again, the relatively good
agreement between the three definitions can be seen for Rig . 0.25. Given the notable
fundamental differences between the two sets of simulations, i.e. the SKIF simulations
are for time evolving homogeneous shear flows while the GVA simulations are for
fully developed wall-bounded turbulent flow, it is encouraging to see similar trends
in the behaviour of all three definitions. This highlights the fact that the fundamental
https://doi.org/10.1017/jfm.2016.340 Published online by Cambridge University Press

physics associated with shear generated turbulence that is locally in a state of quasi-
equilibrium is at play in both of these flows as long as Rig . 0.25, and indicates that
the fluxes are predominantly downgradient (i.e. turbulent) for both flows. On the other
hand, for high Rig , the agreement between the three definitions decreases rather rapidly
(but without oscillations since the data shown are for steady fully developed flow),
especially between the traditional definition Rf and the other two definitions. This is
not surprising given that the production P of k decays as the mean shear rate drops
further away from the wall (i.e. in the inner core of the channel). On the other hand,
in this far-wall region, reversible effects from linear internal waves are dominant and
contaminate the buoyancy flux. Both RIIf and R∗f appear to level out with increasing
Rig , with RIIf somewhat smaller than R∗f , consistent with the trend also seen in figure 1.

3.2. A simple parameterization for the irreversible flux Richardson number R∗f
It is evident from the discussion in § 3.1 that the irreversible flux Richardson number
R∗f is the only definition that is guaranteed to remain positive definite across a broad
range of stratification, a requirement that is necessary in order to be consistent with
798 R1-7
S. K. Venayagamoorthy and J. R. Koseff

0.30

0.25

0.20

0.15

0.10

0.05

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

F IGURE 4. The irreversible flux Richardson number R∗f as a function of the gradient
Richardson number Rig , computed from the DNS data of Shih et al. (2005). The dashed
line shows the fit given by (3.1).

the eddy-viscosity and diffusivity formulations defined in (1.1) which are based on the
gradient-transport hypothesis. Figure 4 shows R∗f as function of Rig as well as a simple
exponential fit that was originally proposed by Karimpour & Venayagamoorthy (2014)
to mimic the popular parameterization for the flux Richardson number proposed by
Mellor & Yamada (1982). The particular form of the exponential function that is
proposed here is given as

R∗f = R∗f ∞ [1 − exp(−γ Rig )], (3.1)

where we set R∗f ∞ = 0.25, and γ is a constant that is set equal to 7. The agreement
between the proposed fit and the data is very good for all Rig . A value of γ = 7 with
a ±1 variation (not shown here) envelops the upper and lower bounds of the data in
the low-Rig regime (Rig . 0.25) very well. Such a parameterization is simple to use
https://doi.org/10.1017/jfm.2016.340 Published online by Cambridge University Press

given that it is based on Rig which is an external parameter that is reasonably easy to
measure in laboratory/field experiments and straightforward to compute in numerical
simulations.

4. Concluding remarks
In this study, our goal was to present a careful comparison of the three definitions
of the flux Richardson number, Rf , RIIf and R∗f , that are commonly used to provide
a measure of the efficiency of the irreversible conversion of turbulent kinetic energy
into background potential energy (for a detailed definition of the background potential
energy, see Lorenz (1955) and Winters et al. (1995)). Using DNS datasets of time
evolving stably stratified homogeneous shear flow as well as channel flow, we
have shown that in the shear dominated flow regime (0 6 Rig . 0.25), all three
formulations are approximately equivalent. This is an important result in that it
allows for the estimation of irreversible mixing directly from flux measurements,
and vice versa for inferring fluxes from indirect estimates of dissipation rates. On
the other hand, our analysis show that the agreement between the three definitions
798 R1-8
Flux Richardson number
deteriorates dramatically as Rig increases above 0.25, with significant oscillations in
both Rf and RIIf in time evolving homogeneous stably stratified shear flows. Both Rf
and RIIf do not separate out the effects of countergradient fluxes which are pervasive
in the buoyancy dominated flow regime. The irreversible flux Richardson number R∗f
is the only formulation that is free from such large oscillations and exhibits a clear
positive definite trend with increasing Rig . We find that a simple exponential fit for
R∗f as a function of Rig mimics the trend in the data remarkably well. As a final note,
it is necessary to evaluate data from flows at much higher Reynolds numbers in order
to ascertain the sensitivity of the trends found in this present study, especially in the
strongly stable flow regime.

Acknowledgements
The authors thank the two anonymous referees for their constructive comments
and recommendations. S.K.V. gratefully acknowledges support from the UPS
Foundation for his sabbatical at Stanford University, during which time this paper was
written. S.K.V. also gratefully acknowledges support from both the National Science
Foundation under grant no. OCE-1151838 and the Office of Naval Research under
grant no. N00014-12-0938 (Scientific Officers: Dr T. Paluszkiewicz and Dr S. Harper).
J.R.K. wishes to acknowledge the generous support from the Stanford Woods Institute
for the Environment.

References
B RITTER , R. E. 1974 An experiment on turbulence in a density stratified fluid. PhD thesis, Monash
University, Victoria, Australia.
E LLISON , T. H. 1957 Turbulent transport of heat and momentum from an infinite rough plane.
J. Fluid Mech. 2, 456–466.
F ERNANDO , H. J. S. 1991 Mixing efficiency in stratified shear flows. Annu. Rev. Fluid Mech. 23,
455–493.
G ARCÍA -V ILLALBA , M. & DEL Á LAMO , J. C. 2011 Turbulence modification by stable stratification
in channel flow. Phys. Fluids 23, 045104.
G REGG , M. C. 1987 Diapycnal mixing in the thermocline. J. Geophys. Res. 92, 5249–5286.
H OLT, S. E., K OSEFF , J. R. & F ERZIGER , J. H. 1992 A numerical study of the evolution and
https://doi.org/10.1017/jfm.2016.340 Published online by Cambridge University Press

structure of homogeneous stably stratified sheared turbulence. J. Fluid Mech. 237, 499–539.
H OWARD , L. N. 1961 Note on a paper of John W. Miles. J. Fluid Mech. 10, 509–512.
I TSWEIRE , E. C., K OSEFF , J. R., B RIGGS , D. A. & F ERZIGER , J. H. 1993 Turbulence in stratified
shear flows: implications for interpreting shear-induced mixing in the ocean. J. Phys. Oceanogr.
23, 1508–1522.
I VEY, G. N. & I MBERGER , J. 1991 On the nature of turbulence in a stratified fluid. Part I: the
energetics of mixing. J. Phys. Oceanogr. 21, 650–658.
I VEY, G. N., W INTERS , K. B. & K OSEFF , J. R. 2008 Density stratification, turbulence, but how
much mixing? Annu. Rev. Fluid Mech. 35, 135–167.
K ARIMPOUR , F. & V ENAYAGAMOORTHY, S. K. 2014 A simple turbulence model for stably stratified
wall-bounded flows. J. Geophys. Res. 119, 870–880.
K ARIMPOUR , F. & V ENAYAGAMOORTHY, S. K. 2015 On turbulent mixing in stably stratified wall-
bounded flows. Phys. Fluids 27, 046603.
L AUNDER , B. E. & S PALDING , D. B. 1972 Mathematical Models of Turbulence. Academic.
L ORENZ , E. N. 1955 Available potential energy and the maintenance of the general circulation.
Tellus 7 (2), 157–167.
L OZOVATSKY, I. & F ERNANDO , H. 2013 Mixing efficiency in natural flows. Phil. Trans. R. Soc.
Lond. A 371 (1982), 20120213.

798 R1-9
S. K. Venayagamoorthy and J. R. Koseff
M ATER , B. D. & V ENAYAGAMOORTHY, S. K. 2014 The quest for an unambiguous parameterization
of mixing efficiency in stably stratified geophysical flows. Geophys. Res. Lett. 41,
doi:10.1002/2014GL060571.
M ELLOR , G. L. & YAMADA , T. 1982 Development of a turbulence closure model for geophysical
fluid problems. Rev. Geophys. Space Phys. 20, 851–875.
M ILES , J. W. 1961 On the stability of heterogeneous shear flows. J. Fluid Mech. 10, 496–508.
O SBORN , T. R. 1980 Estimates of the local rate of vertical diffusion from dissipation measurements.
J. Phys. Oceanogr. 10, 83–89.
O SBORN , T. R. & C OX , C. S. 1972 Oceanic fine structure. Geophys. Astrophys. Fluid Dyn. 3 (1),
321–345.
P ELTIER , W. R. & C AULFIELD , C. P. 2003 Mixing efficiency in stratified shear flows. Annu. Rev.
Fluid Mech. 35, 135–167.
P OPE , S. B. 2000 Turbulent Flows. Cambridge University Press.
ROGERS , M. M., M ANSOUR , N. N. & R EYNOLDS , W. C. 1989 An algebraic model for the turbulent
flux of a passive scalar. J. Fluid Mech. 203, 77–101.
ROHR , J. J., I TSWEIRE , E. C., H ELLAND , K. N. & ATTA , C. W. V. 1988 Growth and decay of
turbulence in a stably stratified shear flow. J. Fluid Mech. 195, 77–111.
S HIH , L. H., K OSEFF , J. R., I VEY, G. N. & F ERZIGER , J. H. 2005 Parameterization of turbulent
fluxes and scales using homogeneous sheared stably stratified turbulence simulations. J. Fluid
Mech. 525, 193–214.
S TRETCH , D. D., ROTTMAN , J. W., V ENAYAGAMOORTHY, S. K., N OMURA , K. K. & R EHMANN ,
C. R. 2010 Mixing efficiency in decaying stably stratified turbulence. Dyn. Atmos. Oceans
49, 25–36.
V ENAYAGAMOORTHY, S. K. & S TRETCH , D. D. 2010 On the turbulent Prandtl number in
homogeneous stably stratified turbulence. J. Fluid Mech. 644, 359–369.
W INTERS , K. B., L OMBARD , P. N., R ILEY, J. J. & D’A SARO , E. A. 1995 Available potential
energy and mixing in density-stratified fluids. J. Fluid Mech. 289, 115–128.
https://doi.org/10.1017/jfm.2016.340 Published online by Cambridge University Press

798 R1-10

You might also like