The Myth of URANS

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Journal of Turbulence

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/tjot20

The myth of URANS

Daniel M. Israel

To cite this article: Daniel M. Israel (2023) The myth of URANS, Journal of Turbulence, 24:8,
367-392, DOI: 10.1080/14685248.2023.2225140
To link to this article: https://doi.org/10.1080/14685248.2023.2225140

© 2023 The Author(s). Published by Informa


UK Limited, trading as Taylor & Francis
Group

Published online: 19 Jun 2023.

Submit your article to this journal

Article views: 916

View related articles

View Crossmark data

Citing articles: 1 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tjot20
JOURNAL OF TURBULENCE
2023, VOL. 24, NO. 8, 367–392
https://doi.org/10.1080/14685248.2023.2225140

The myth of URANS


Daniel M. Israel

Continuum Models and Numerical Methods Group, Los Alamos National Laboratory, Los Alamos, NM, USA

ABSTRACT ARTICLE HISTORY


Since the 1990s, RANS practitioners have observed spontaneous Received 6 December 2022
unsteadiness in RANS simulations. Some have suggested deliber- Accepted 1 May 2023
ately using this as a method of resolving large turbulent structures. KEYWORDS
However, to date, no one has produced a theoretical justification for Reynolds averaged
this unsteady RANS (URANS) approach. Here, we extend the dynam- Navier–Stokes; unsteady
ical system fixed point analysis to create a theoretical model for RANS; homogeneous
URANS dynamics. The results are compared to URANS simulations for turbulence
homogeneous isotropic decaying turbulence. The model shows that
URANS can predict incorrect decay rates and that the solution tends
towards steady RANS over time. Similar analysis for forced turbulence
shows a fixed modelled energy of about 30% of total energy, regard-
less of the model parameters. The same analysis can be used to show
how hybrid type models can begin to address these issues.

1. Introduction
When Reynolds [1] introduced the averaged equations that bear his name, he posited an
average that was local over a small area of space or interval of time. Inspired by an analogy
to kinetic theory, he assumed that

ūu = 0, (1)

which eliminates the cross-stress terms in the averaged equations. For kinetic theory, this
is an excellent approximation: there is an extreme scale discrepancy between the mean-
free path that characterises molecular motions and the scales of interest in a continuum
representation of a flow, at least for most flows of interest. For turbulence, however, which
is characterised by the broad spectrum of the energy cascade, and assuming a local space
or time average as Reynolds did, this assumption is wrong.
However, there are numerous other averages for which Equation (1) does apply, such
as a homogeneous spatial average, infinite time average, or ensemble average. Subsequent
literature typically speaks of a ‘Reynolds average’, if the averaging operator has the property

f ḡ = f̄ ḡ, (2)

CONTACT Daniel M. Israel dmi1@lanl.gov Continuum Models and Numerical Methods Group, Los Alamos
National Laboratory, P.O. Box 1663, Mailstop F644, Los Alamos, NM 87545 USA
© 2023 The Author(s). Published by Informa UK Limited, trading as Taylor & Francis Group
This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivatives License
(http://creativecommons.org/licenses/by-nc-nd/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium,
provided the original work is properly cited, and is not altered, transformed, or built upon in any way. The terms on which this article has been
published allow the posting of the Accepted Manuscript in a repository by the author(s) or with their consent.
368 D. M. ISRAEL

which is a generalisation of Equation (1). Ironically, that makes the local spatial average
posited in Reynolds’s original paper not a Reynolds average, but as Reynolds assumed
Equation (2) in analogy with kinetic theory, where it does hold due to the extreme scale sep-
aration present, the equations derived in Reynolds’s paper are consistent with a Reynolds
average so defined.
It is evident that averages of this kind are expected to remove all turbulent scales of
motion, including the very largest eddies. The earliest attempts to close these equations,
such as Taylor [2], Prandtl [3], assume that the turbulence is characterised by a single set of
scales, those corresponding to the largest, energy-containing eddies. This class of models,
which comes to be known as RANS (Reynolds-averaged Navier–Stokes), continues to rely
on the single scaling assumption. This includes k − ε, k − ω, and Reynolds-stress models
(RSM), all of which include one scaling equation which is used to obtain all the remaining
unclosed quantities by a dimensional scaling. All the proposed scaling quantities (dissipa-
tion, ε, inverse time scale, ω, or length scale, among others) are assumed to be algebraically
related.
With the first availability of larger computers in the 1960s, starting with the pioneering
work of Smagorinsky [4], Lilly [5], and Deardorff [6], turbulence modelling returned to
the idea of a local spatial average over regions significantly smaller than the largest tur-
bulent eddies. This class of methods is called large-eddy simulation (LES). The averaging
region is often identified with the computational grid cell. For example Lilly [5] and Dear-
dorff [6] identify the averaging operator used in deriving the LES governing equations with
the average over the computational cells used in computing the solution. However, they are
mathematically distinct, and this conflation continues to cause confusion.
The RANS literature often derives the averaged equations by invoking a long or infinite-
time average, or an ensemble average, whereas LES papers typically formulate the filtered
equations in terms of a local spatial filter. This has led to a widespread misunderstand-
ing that the choice of filter used to derive the equations is what distinguishes RANS and
LES. However, as Germano [7] points out, by writing the equations in terms of generalised
moments, the moment equations are identical for any average, provided it is linear, and
commutes with differentiation. So the choice of average used to derive the exact moment
equations cannot be what makes one model RANS and another LES. Instead, it must be the
choice of model used to close the equations. Indeed, it is obvious that this must be the case,
since two authors who derive the equations using different filters, but use the same closure
model, will obtain identical results, whereas two authors who derive the equations using
the same filter, but choose different closure models, will obtain different results. In other
words, the average is implicitly defined by the closure, rather than the opposite. This point
is made very clearly by Pope [8] for the specific case of the Smagorinsky model, and what
he terms the implied Smagorinsky filter, but the argument generalises to any turbulence
model.
In particular, while RANS models rely on the single scaling assumption described above,
LES models invoke a cutoff scale that defines the largest scale at which the model acts. This
is typically a length scale, and usually is related to the grid scale to assure maximum use of
the available resolution without too large a discretisation error.
It has long been known that RANS models can be used to compute time-dependent
flows in cases where the model can be considered a homogeneous spatial average, for exam-
ple, homogeneous isotropic decaying turbulence or the temporal spreading of a mixing
JOURNAL OF TURBULENCE 369

layer. For these examples, the RANS equations reduce to ordinary differential equations
in the first case and partial-differential equations in time and a single space dimension
in the second. Clearly, no turbulent structures are explicitly resolved. Alternatively, one
might consider a flow in which the time scale of interest is very long compared to the tur-
bulent time scale, for example, the slow spread of a contaminant plume. In addition, it has
been common practice to solve for the time-steady RANS solution to a problem by using
a time-dependent code and iterating in time until a steady-state is achieved.
As early as 1985, Celenligil and Mellor [9] observed unsteadiness in the simulation
of a backward-facing step using an RSM. Lasher and Taulbee [10], observing a similar
unsteady solution, raised theoretical concerns with computing an unsteady solution using
a steady model, and pointed out the need for further research. Claus et al. [11] equated
the solution of the Navier–Stokes equations using a finite-volume algorithm to solving the
box-filtered equations and suggested ‘this permits the use of a turbulence model that is sim-
ply a time-dependent form of the commonly used Reynolds-averaged k − ε model ’. They
described this as an LES model, however, their paper gave no indication why they expect a
model which was developed and successfully calibrated for steady solutions [12,13] should
work as an unsteady subfilter stress model, or even give rise to unsteadiness at all. They
applied the model to a forced shear layer; it is unclear to what extent the unsteadiness
in the resulting simulations is due to the convective instability versus the forcing. The
simulations reproduced some qualitative aspects of the flow, but significant questions
remain, including to what extent the results (which are two-dimensional) are still grid
dependent.
A series of studies by Rodi et al. [14–16] considered the vortex shedding behind a square
cylinder. In their presentation, the average is described as an ensemble- or phase-average,
which is closed by a statistical model intended to represent the stochastic turbulent fluc-
tuations, and the mean and periodic components are resolved. Again, without further
justification, they adopt as their statistical model several variations of the k − ε model [13]
and the RSM of Launder et al. [17]. For the k − ε model with wall functions, they found
that the emergence of an unsteady solution was dependent on the numerics. In general,
they had reasonable agreement with experimental data, although there was some uncer-
tainty associated with how to set certain initial conditions which are not known from
experiments.
Similar simulations from Chalmers University considered vortex shedding behind a
flameholder [18] and in the wake of a film-cooling slot [19]. They also used the standard
k − ε model (an algebraic stress model was also used in [19]). For the flameholder without
afterbody, they obtained good agreement for the shedding frequency, as well as mean pro-
files of velocity and turbulent kinetic energy, with the model contributing about 50% of the
energy. However, while the experiments demonstrated a change in shedding frequency in
the presence of an afterbody, the simulations could not capture this effect [18]. For the film-
cooling slot, when the tip-to-slot thickness ratio was small, the resulting calculated flow was
steady, and both the k − ε and algebraic-stress models captured the mean profiles well.
With a larger tip thickness, the vortex shedding becomes a more important mechanism,
and both models produced an unsteady solution, but the agreement with data was much
poorer. Durbin [20] looked at the same flameholder geometry as [18] using a k − ε − v2
model, and found a good match for the time averaged velocity profiles. However, he also
reported good results for time-steady simulations of flow over a backward-facing step
370 D. M. ISRAEL

and a separated diffuser using the same model, whereas preliminary computations of a
square cylinder and a thick splitter plate were unsteady. He attributed unsteadiness to the
model representing an ensemble, rather than a time, average, but gave no clear reason why
unsteadiness was present for some geometries but not others.
Numerous subsequent researchers have used RANS models to develop unsteady struc-
tures deliberately, reporting mixed success. For example, [21] showed URANS was unable
to reproduce the back pressure on a cylinder, as compared to detached eddy simulation.
For flow behind a cube, [22] reports good agreement for the mean-velocity gradients. The
recirculation lengths are much better than steady RANS, but not as good as conventional
LES, and the upstream separation length is closer to that for steady RANS. Simulations
of a square cylinder with rounded corners [23] showed similar results for URANS and
detached-eddy simulation (DES), with both performing much better than simulations
without a turbulence model. In a comparison of two- and three-dimensional URANS, [24]
indicated that the latter can lead to significant improvements, but results can be dependent
in unpredictable ways on the choice of model and spanwise domain size. Other authors
who simply identified a conventional RANS model with the solution to a phase-average
rather than a time average include [25,26].
Fadai-Ghotbi et al. [27] considered a backward facing step using a standard k − ε,
LRR-IP, and an elliptic-blending RSM, computing in two dimensions. The k − ε model
produced a steady result. For the RSM, the unsteadiness depended on the numerical
method and the resolution. For the more dissipative upwind scheme, the solutions were
steady for all of the models, whereas, with a central-difference scheme, the two RSMs
gave unsteady solutions. However, they observed that the amplitude of the unsteadiness
decreased with increasing mesh resolution. They concluded that the unsteadiness was trig-
gered by a numerical instability in the shear layer at the corner. Overall, they found that
the shedding frequency was accurately predicted by the URANS, because it is controlled by
the linear instabilities of the mean velocity gradient, but the amplitudes were not physical.
Based on a careful study of shedding behind a circular cylinder, Palkin et al. [28] con-
cluded that, for high-Reynolds number industrial flows, URANS using an RSM can be
reliably employed. This study used comparisons between LES and RANS using two RANS
models, a two-equation elliptical relaxation linear eddy-viscosity model and an RSM, and
showed good agreement for the RSM in their higher-Reynolds number case. However, at
lower Reynolds number, while shedding frequencies were well predicted, amplitudes were
grossly over-predicted. Furthermore, although in presenting the models as closures for the
phase-averaged quantities they assumed a large scale-separation between the periodic and
stochastic components in the triple-decomposition, their spectra show that this is not the
case. Finally, in their conclusions they noted: ‘The success in general is unpredictable . . .
in a sense that it is difficult to know a priori whether the method will return reliable and
realistic results in unknown situations’.
A more recent study comparing RANS and hybrid models suggests that many of the
discrepancies in the preceding results can be explained in terms of how well the develop-
ment of Kelvin–Helmholtz rollers in the turbulent shear layer is resolved [29]. Since the
disturbance growth can be damped by either the numerics or the modelled dissipation, a
full understanding must take into account the numerical method, grid resolution, choice
of model, and how the turbulence quantities are initialised.
JOURNAL OF TURBULENCE 371

The results of these and other studies sometimes look promising, but raise numerous
questions. First, and most fundamentally, it is not clear why some models, geometries,
and initial conditions lead to unsteady solutions, while others relax to a steady solution.
Second, the reported results are mixed; some authors showing relatively good agreement
with experimental data, and others quite poor. This opens the question whether the more
successful comparisons are purely fortuitous. At the very least, our inability to predict, a
priori, when the method will work makes it impossible to rely on for engineering pre-
diction. Finally, some quantities are easier to predict than others. Qualitative agreement
between visualisations of large-scale features is not surprising, since the mechanisms that
support them are often only weakly dependent on Reynolds number.
Spalart [30] and Travin et al. [31] present systematic critiques of what they refer to as
the unsteady RANS (URANS) approach. There is some variation in terminology between
various authors. Consistent with the definitions above, we will follow the distinction made
by Travin et al. [31] that LES refers to a closure model which includes the filter width as a
model parameter,1 whereas RANS does not. We can also distinguish between VLES, which
is simply an LES model that resolves a smaller portion of the spectrum than regular LES,
and URANS, which is just using a traditional, unmodified RANS model but allowing the
solution to become unsteady.2
Although in widespread use, a physical or mathematical justification for URANS is
missing. (‘URANS does not have a rational theoretical background/justification’ [31].)
The underlying models typically used were all developed for steady flows, with all scales
modelled. The reason why unsteadiness arises in certain cases, and not in others, is not
understood, other than a general idea that it is related to how unstable the flow is, and
how dissipative is the model. Physical plausibility arguments for URANS seem to require a
spectral gap between the large structures and the incoherent turbulent fluctuations, which
is not observed in experiment [30,31].
The inability to specify exactly what the URANS solution is supposed to represent makes
it impossible to determine if it is correct or not. Several authors have invoked a triple
decomposition or a phase average [16,18,27,28]. But this is an ex post facto reinterpreta-
tion of models that were originally derived without any accounting for large-scale resolved
structures. The models are the same whether the averages in the equations are identified as
ensemble averages or phase averages. In other words, there is no formal mathematical jus-
tification for interpreting the results this way. Furthermore, Spalart [30] points out that the
phase-averaged structure is very different from the typical exact structure, and the resolved
field is observed to include ‘incoherent’ motions.
An alternative to URANS is to find a rational procedure for introducing an explicit
decomposition parameter into a conventional RANS model. The flow-simulation method-
ology (FSM [32–34]), partially-averaged Navier–Stokes (PANS [35–37]) and partially-
integrated transport model (PITM [38,39]) are all examples of this approach. The
emergence of these methods, which can easily be implemented in existing RANS codes,
has reduced the prevalence of URANS, but has not completely replaced it.
It is the purely anecdotal experience of this author, speaking with colleagues, that they
fall into two camps: one that views URANS as self-evidently an ill-constructed model, and
one that sees it as equally self-evidently valid. Neither side sees it necessary to expend much
effort trying to either prove or disprove it. This, therefore, is the myth alluded to in the title:
372 D. M. ISRAEL

that merely by computing with a conventional RANS model in a time dependent fashion,
reasonable results can be unambiguously obtained.
Without a better theoretical underpinning for the URANS approach, it will remain in
scientific limbo. A clear mathematical framework, on the other hand, can also help us better
understand URANS in comparison with the other available modelling approaches. Travin
et al. [31] writes, ‘Much work and discussion is needed to determine whether this situation
reflects a congenital flaw, or is hiding a “golden opportunity”’. The goal of this paper is to
present a first step in that work and discussion.
There are two primary questions to ask about URANS. First, what is the mechanism that
gives rise to unsteadiness, and when does it occur? Second, in the absence of a parameter to
set the scale for the energy decomposition, where does that scale come from and what value
does it take? We will focus on the second question, by considering a simple test problem,
and constructing an analytic model of how the energy decomposition evolves.
The test problem chosen is homogeneous isotropic turbulence, both decaying and
forced. This problem is not one in which URANS is typically employed and does not exhibit
the large coherent structures seen in such geometries as bluff body wakes. In that sense, it
is a particularly stringent test case. Nevertheless, it was selected for two reasons. First, it is
amenable to an analytic theory which can help guide our intuitive understanding. Second,
it represents the extreme in terms of lack of scale separation; if URANS can still work in
this case, it should be able to work in cases with dynamically significant large structures.

2. The conventional RANS model


We start by reviewing the conventional RANS model. In the following discussion, we
restrict ourselves to the standard k − ε model [13], although the same analysis could easily
be extended to other RANS models. We decompose the velocity and pressure fields into an
averaged and a fluctuating part,

ui = ūi + ui (3a)



p = p̄ + p . (3b)

The governing equations are

∂ ūi
=0 (4a)
∂xi
∂ ūi ∂ ūi 1 ∂ p̄ ∂Rij
+ ūj =− − + ν∇ 2 ūi (4b)
∂t ∂xj ρ ∂xi ∂xj
  
∂k ∂k ∂ νT ∂k
+ ūj =P −ε+ ν+ (4c)
∂t ∂xj ∂xj Prk ∂xj
  
∂ε ∂ε ε ∂ νT ∂ε
+ ūj = (Cε1 P − Cε2 ε) + ν+ . (4d)
∂t ∂xj k ∂xj Prε ∂xj

The Reynolds stress is defined as

Rij = ui uj , (5)


JOURNAL OF TURBULENCE 373

and modelled with a linear eddy viscosity


2
Rij = −2νT S̄ij + kδij , (6)
3
with
 
1 ∂ ūi ∂ ūj
S̄ij = + . (7)
2 ∂xj ∂xi
The eddy viscosity is
k2
νT = C μ , (8)
ε
and the production of turbulent kinetic energy is
∂ ūi
P =− Rij = 2νT S̄ij S̄ij . (9)
∂xj
The over-bar represents an average, which is often taken to be a homogenous spatial aver-
age, for problems with one or more homogeneous spatial dimensions, or a time average,
for steady problems. For more general problems, it is common to invoke an ensemble aver-
age. These averages are often equivalent, assuming ergodicity holds. The turbulent kinetic
energy,
1
k = ui ui , (10)
2
and the dissipation rate,

∂ui ∂ui
ε=ν , (11)
∂xj ∂xj
are then the energy and dissipation in the scales of motion too small to be resolved in the
ūi velocity field.
In fact, the exact averaged equations, which these equations are a model for, are iden-
tical, regardless of which average is invoked. This implies that the average that is obtained
from an actual simulation is a consequence of the specific model closure used, and is inde-
pendent of the average specified by the author, if any. In the case of the k − ε model, the
average is implied by the modelling assumption that the eddy viscosity can be scaled with
a velocity characteristic of the entire turbulent motion, k1/2 , and a length scale
k3/2
L∼ ,
ε
as well as the rescalings that go into modelling the dissipation rate. Consistent with our def-
initions in Section 1, the RANS has no explicit parameter defining a partition of resolved
and modelled scales. An alternative closure approach would be to explicitly introduce such
a parameter to obtain an LES model. For example, if we impose a length scale for comput-
ing the eddy viscosity, L ∼ CS x, then the equations are immediately closed without the
need for a dissipation rate equation (4d), and we recover the one-equation LES model of
Schumann [40] and Yoshizawa [41].
374 D. M. ISRAEL

Particular care must be given to initial and boundary conditions. Equations (4a) allow
for injecting both a modelled and a resolved component of turbulent fluctuations through
the initial and boundary conditions. There is nothing in the equations, however, which
ensures that this injection is consistent with the energy decomposition implied by the
model. In fact, it is fair to ask whether the possibility of forcing the model in an incon-
sistent manner constitutes a defect in the model, or just a caveat on how we treat initial
and boundary conditions.
Applying Equation (4a) to homogeneous isotropic decaying turbulence, using the con-
ventional RANS approach, we assume that the average removes all the turbulent scales.
Since the flow is statistically homogeneous, all the average quantities are uniform in space,
and the equations reduce to ordinary differential equations. To indicate this, we will replace
the over-bar with an angle bracket. This could represent a spatial average over all of space,
or an ergodically equivalent, suitably constructed, ensemble average. The average velocities
are identically zero, ui  = 0, and the k − ε equations reduce to a set of ordinary differential
equations,
d k
= − ε (12a)
dt
d ε ε2
= −Cε2 . (12b)
dt k
The solution to this problem is well known to be
 −n
t
k = k0 , (13)
t0
where the decay rate is
1
n= . (14)
Cε2 − 1
For this simple RANS model, the decay rate is a constant. Data from experiments and
simulations suggest that this may not be quite right, but for a particular experiment, over
a wide range of Reynolds number, it is a reasonable approximation [42,43].

3. Unsteady RANS
We can also try to simulate this problem using a URANS. This means we use the same k − ε
RANS equations (4), but now we include all the unsteady and locally varying terms. For this
problem, unsteadiness will not arise spontaneously, so we need to introduce it through the
choice of initial conditions. We wish do to so in a way that is consistent with the idea that the
initial conditions represent some suitably averaged real turbulent state. Conceptually, we
can achieve this by taking a fully resolved turbulent field (perhaps from direct-numerical
simulation) and filtering it using any arbitrary low-pass filter characterised by a specified
length-scale. Using this filter, we can obtain filtered velocity fields. The turbulence quanti-
ties can be obtained using Equations (10) and (11). The initial conditions obtained this way
will depend on both the length scale and the filter shape (Gaussian, top-hat, spectral cut-
off, etc.). The effect of the shape is likely to be small, but may lead to a short initial transient.
Appendix 2 provides additional detail of how this procedure was done in practice.
JOURNAL OF TURBULENCE 375

The governing equations are now


∂ ūi
=0 (15a)
∂xi
  
∂ ūi ∂ ūi 1 ∂ p̄ ∂  <
 ∂ ūi ∂ ūj
+ ūj =− + ν + νT + (15b)
∂t ∂xj ρ ∂xi ∂xj ∂xj ∂xi
 <  
∂k< ∂k< ∂ νT ∂k<
+ ūj = P< − ε< + ν+ < (15c)
∂t ∂xj ∂xj Prk ∂xj
  
∂ε< ∂ε< ε<  < <
 ∂ νT< ∂ε<
+ ūj = C P< − Cε2 ε< + ν+ < , (15d)
∂t ∂xj k< ε1 ∂xj Prε ∂xj
with

P< = 2νT< S̄ij S̄ij , (15e)

and
k2<
νT< = Cμ . (15f)
ε<
Note, these equations are formally identical to the RANS equations (4), however, we have
introduced super- or subscript less-than decorations to various quantities to emphasise that
they now represent subfilter quantities. Both the velocities and the turbulence quantities
now vary in both time and space.
The bar now represents a local filter, which does not remove all turbulent scales, and the
term average, denoted by angle brackets will be reserved for the homogeneous average over
the entire box. It is important to note that, as mentioned above, the filter corresponding to
a specific model is implicitly defined by the model, and is therefore not known in a closed
form that can be used to compute the initial conditions. We will assume that the effect of
using slightly different low-pass filters leads to, at most, minimal, short-lived transients as
the flow adjusts.
Unlike LES models, which have an explicit cut-off length scale specified, when using
a conventional RANS model to try to compute unsteadiness, the model does not include
a parameter that controls which scales, or how much of the energy, is to be captured by
the model. This makes it difficult to say what constitutes the correct answer. Instead we
will focus on two questions. How does the partition of energy between resolved and model
scales evolve? And, how well does the URANS predict the overall decay rate? The first
question addresses how useful the model is: given that we do not control the energy parti-
tion, does it behave in a manner that produces an informative result? The second question
is more about correctness. There may not be a clear right answer for each component,
resolved and modelled, of the energy, but the total energy should give something like a
power-law decay, as observed in experiments, with at most a slowly varying decay rate. In
particular, if the energy partition changes, it should do so in a way that does not degrade
the solution for the total energy.
To explore these questions, it would be nice if we could reduce Equations (15) to a set
of ordinary differential equations. We could then employ the dynamical systems and fixed
point analysis first introduced by Speziale and Mhuiris [44] in the steady RANS context
376 D. M. ISRAEL

and applied to URANS by Girimaji et al. [35]. In their approach, they omitted the unsteady
transport term as not contributing to the global energy balance. We take a slightly different
approach and formally average the unsteady equations. To do this, we first introduce the
decomposition

k = k>  + k<  ,

where
1
k< = (ui ui − ūi ūi ) (16a)
2
1
k> = (ūi ūi − ūi  ūi ) (16b)
2
are the energy of the subfilter and resolved scales, respectively. This decomposition requires
the assumption that f̄  ≈ f . This can be demonstrated explicitly for certain filters and
averages. Similarly, for the dissipation,
 
ε< = 2ν Sij Sij − S̄ij S̄ij (17a)
   
ε> = 2ν S̄ij S̄ij − S̄ij S̄ij . (17b)

Taking the average of Equation (15c) and using homogeneity, the equation for the
subfilter energy is exactly

∂ k< 
= P<  − ε<  . (18)
∂t
The equation for the average resolved energy can be derived by multiplying Equation (15b)
by ūi , and averaging,

∂ k> 
= − P<  − ε>  . (19)
∂t
The subfilter dissipation Equation (15d), when averaged, yields

∂ ε<  < ε< < ε<


2
= Cε1 P< − Cε2 . (20)
∂t k< k<

This is an exact equation for how the subfilter dissipation rate will evolve under the action
of the RANS model given by Equation (4a). To solve this equation, an additional level of
closure modelling is required, since the two averaged terms on the right-hand side are not
closed. That is, this equation includes a closure model relative to the bar filter, but still
requires closing relative to the bracket filter. The simplest model for Equation (20) is

ε<
2 ε< 2
≈ (21)
k< k< 
ε< ε< 
P< ≈ P<  . (22)
k< k< 
JOURNAL OF TURBULENCE 377

This is correct to second order in the fluctuations [45]. For the remaining unclosed
quantities, we use

P<  ≈ νT  ω̄2 (23a)
 2
ε>  = ν ω̄ , (23b)

where ωi is the vorticity. We also decompose the average eddy viscosity as

k< 2
νT  = Cμ . (24)
ε< 
Note that Equation (23b) is exact, whereas Equation (23a) neglects the spatial variation of
the eddy viscosity.
We can now close the three equations (18), (19) and (20) by adding an evolution
equation for filtered enstrophy magnitude, ω̄2 . To obtain this we start by noting that the
dissipation equation (12b) is essentially an enstrophy equation,
  2
d ω2 ν ω2
= −Cε2 . (25)
dt k
If we assume that the dynamics of a URANS simulation can be modelled as a Navier–Stokes
simulation with a increased viscosity of νT , rather than ν, then a plausible model for the
enstrophy of the resolved scales is
  22
∂ ω̄2 > νT  ω̄
= −Cε2 . (26)
∂t k> 
(A similar argument can be found in [8] for deriving an estimate of the filtered spectrum
produced by an LES model.)
The coefficient Cε2> plays the same role for the enstrophy dissipation of the filtered scales
<
as the Cε2 plays for the subfilter scales in the original RANS model and is the only parame-
ter of the reduced order model. The parameters Cε1 < and C< are not tunable parameters of
ε2
the reduced order model, they must be set to whatever values are used in the URANS sim-
ulations we are comparing to. For the results presented here, the value used is the standard
Cε2 parameter setting, Cε2 > = 1.92.

The complete set of modelled equations is now


∂ k> 
= − P<  − ε>  (27a)
∂t
  2
∂ ω̄2 > νT  ω̄2
= −Cε2 (27b)
∂t k> 
∂ <
k
= P<  − ε<  (27c)
∂t
∂ <
ε ε<   < <

= Cε1 P<  − Cε2 ε<  . (27d)
∂t k< 
The first thing to note about these equations is that if we sum Equations (27a) and (27c),
we recover the exact total energy Equation (12a). (This is, in fact, a property of the exact
378 D. M. ISRAEL

unclosed equations, so will hold for any closure model.) It is evident that the subfilter pro-
duction term, P<  represents the transfer of energy from the resolved to the subfilter
scales. As such, it is the primary term which sets the decomposition scale. This point is
widely misunderstood and bears emphasis. The decomposition into resolved and subfilter
components is not unique, and the partition of energy is set implicitly by the model through
the action of P< .
Summing Equation (27d) and viscosity times Equation (27b) does not recover the total
dissipation equation (12b), however. Instead, using Equations (23) to eliminate ν and νT ,
we obtain
 
∂ε < ε<  > ε>  < ε< 
2
= Cε1 − Cε2 P<  − Cε2 . (28)
∂t k<  k>  k< 

If we assume that the corresponding terms play similar roles in the dissipation equations as
they do in the kinetic energy equations, then the first two terms represent the dissipation
of enstrophy out of the large scales, and into the small scales, and should therefore can-
cel. The last term should represent the dissipation of small-scale dissipation and should
equal the right-hand side of Equation (12b). That this is not the case allows our model
for the URANS to explore deviations from the RANS power law decay. This is in distinc-
tion from the analysis of Girimaji et al. [35], which assumes the total dissipation still obeys
Equation (12b).
Rather than analysing the full set of Equations (27), we can construct a reduced order
model in terms of the primary quantities of interest. The first of these is the energy
partition, defined as

k< 
f = . (29a)
k
Using the product rule and substituting Equations (27c) and (12a), we can write an
evolution equation for this quantity,
 
df 1 d k<  dk
= 2 k − k< 
dt k dt dt
  
ε<  P<  k<  ε> 
= f −1+ +1 .
k<  ε<  k ε< 

Defining two other reduced-order quantities, the subfilter production to dissipation ratio,
and the subfilter Reynolds number,

P< 
g= (29b)
ε< 
νT 
R= , (29c)
ν
and a non-dimensional time
dt k< 

= , (30)
dt ε< 
JOURNAL OF TURBULENCE 379

results in the final equation


df  

= R−1 fg + f + g − 1 f . (31a)
dt
Similarly, we can write
  
dg d P<  d k< 2 ω̄2
= = Cμ ,
dt dt ε<  dt ε< 2

and, using Equations (27b), (27c) and (27d), obtain


 
dg  <  > f
 <  2
= 2 C ε2 − 1 g − Cε2 + 2 Cε1 − 1 g . (31b)
dt ∗ 1−f
Finally, for R, we have
 
dR d νT  d k< 2
= = Cμ .
dt dt ν dt ν ε< 

Again, using Equations (27c) and (27d), yields


dR  <
 <


= 2 − Cε1 g + Cε2 − 2 R. (31c)
dt
The energy partition, f, is the ratio of the subfiliter energy to the total energy. It repre-
sents the fraction of the energy spectrum which is being modelled. In the limit f = 1, the
simulation is effectively a conventional RANS, with all scales modelled. The other limit,
f = 0, is a DNS, with no model contribution. The variable g is the ratio of the subfilter
production to the subfilter dissipation, that is, the rate at which energy is moving from
resolved to modelled scales divided by the rate in which modelled scale energy is being
dissipated by viscosity. Finally, R, although it can be written in a form reminiscent of the
turbulent Reynolds number of the subfilter scales, is probably best understood as the nor-
malised eddy viscosity due to the subfilter scales. That is, R is the contribution to the total
effective viscosity due to the modelled scales.
This model has a singularity when the subfilter Reynolds number goes to zero, but for
even moderate Reynolds number, the singular term is negligible, and the f and g equations
evolve independently of R. We can better understand the behaviour of R by noting that
g+R 2
R = Cμ f Ret ,
R
where the turbulent Reynolds number is defined as

k2
Ret = . (32)
νε
Therefore the singular term behaves as
R g −1
R−1 fg = Cμ
−1
Re .
g+Rf t
380 D. M. ISRAEL

This term will blow up in two situations. The first is when Ret → 0, that is, when the tur-
bulence is almost completely dissipated. We can neglect this case as uninteresting to our
current investigation. The second is when f → 0, which is the DNS limit, when the flow is
almost completely resolved. For this latter situation, for this term to be O(1), we must have
f ∼ Re−1
t . This case can also be ignored, since if f was that small, we would just run DNS
instead of URANS. Consequently, the reduced-order model can be well approximated as
df  

= f +g−1 f (33a)
dt
 
dg  <  > f
 <  2
= 2 Cε2 − 1 g − Cε2 + 2 Cε1 − 1 g . (33b)
dt ∗ 1−f
The same equations can be obtained directly by neglecting the viscous dissipation in the
resolved scales, ε> , in Equation (27a).
This model has four fixed points,

f =0 g=0 (34a)
<
Cε2 −1
f =0 g= < (34b)
Cε1 −1
f =1 g=0 (34c)
< − 2C<
2Cε1 <
2Cε2 >
− Cε2 −2
ε2
f = < − C> − 2 g= < > . (34d)
2Cε1 ε2 2Cε1 − Cε2 −2
The last one is an attractor that governs the behaviour of the system. As can be seen in
Figure 1, all trajectories tend towards this fixed point, which is very near the RANS limit.
For comparison, Figure 1 also shows several real trajectories from numerical experi-
ments, in red. To obtain these, a fully resolved DNS simulation for forced homogeneous
turbulence was run until it reached equilibrium. The velocity fields obtained were filtered
(as described above) to obtain initial conditions for the URANS. Each trajectory was cre-
ated using a different filter width to generate the initial condition. The trajectory was then
computed using a URANS. Details of the simulations can be found Appendix 2.

Figure 1. Phase space evolution of Equation (33a). The grey trajectories are the reduced order model,
with the fixed points in black. Red lines are trajectories from simulation data (with points indicating the
beginning of each trajectory). The green shaded region is where n = 1.15−1.45. The dashed curve is
Equation (A12).
JOURNAL OF TURBULENCE 381

What the figure clearly shows is that, regardless of the initial condition, the URANS
evolves towards something very like a steady RANS, with the resolved fluctuations decay-
ing away. The model suggests that the URANS will not go all the way to the RANS limit of
f = 1, however, the data is not sufficient to determine if this is the case. What is clear is that
the user has no control over the range of scales which are physically resolved. However, it
should be noted that the rate at which the URANS relaxes towards RANS is extremely slow
in turbulence time scales, which may be one reason why some URANS practitioners are
left with the impression that the initial choice of energy partition persists.
Having considered the evolution of the energy partition, we now examine the decay rate
of the total turbulent kinetic energy. For a flow in which the decay rate may be changing,
there are several ways to measure this [46]. We choose
  −1
d k
n= , (35)
dt ε
which has the advantage that it can be computed from Equations (27) in terms of our
reduced order model,
f
n= < <g − f . (36)
Cε2 − Cε1
On our trajectory plot, curves of constant n are straight lines through the point f = 0,
g = Cε2 < /C< . The green shaded region in Figure 1 is where the decay matches the observed
ε1
range of experimental data, n = 1.15−1.45.
It is clear that the shaded region, where the decay rate matches experiment, occupies
only a small fraction of phase space. Whether this is a problem for the URANS approach
in practice depends largely on how much time the URANS spends outside this small region.
The plotted trajectories suggest that the URANS does evolve to this region, given sufficient
time. Another question might be, where do typical initial conditions fall relative to this
region. For example, given a local spatial filter, characterised by a cutoff length-scale, we
could filter DNS data at various cutoffs, to obtain an initial condition for each one. Each
initial condition would define one point in phase space, and the points for all possible filter
widths would define a curve. We can approximate that curve using a stick-spectrum anal-
ysis, to obtain the dashed curve, Equation (A12), shown in Figure 1. (The details of the
derivation are given in Appendix 1). From this analysis, it appears that the initial condi-
tions thus produced fall well below the shaded region, meaning there will be a significant
adjustment before the correct energy decay rate is observed. The red trajectories, on the
other hand, start well above (at least for small f ), the difference probably being a result of
the specific way they are generated, starting with a forced turbulence and only later being
allowed to decay, as described in Appendix 2.
An interesting feature of the reduced-order model is that if we use Equations (34d)
and (36) to compute the decay rate at the fixed point, we get
2
n= >. (37)
Cε2
In other words, the decay rate is independent of the underlying RANS model, and is a
function of the dynamics of the resolved scales alone. The model coefficient we chose pre-
dicts a decay rate of n ≈ 1.02, noticeably lower than might be expected from experiments.
However, from Equation (36) we see that changing Cε2 > does not actually change the decay
382 D. M. ISRAEL

Figure 2. The measured decay rate compared versus the decay rate predicted by the model
(Equation 36) for the URANS simulations [47] shown by the trajectories in Figure 1. The cases correspond
to larger initial f in the direction of the arrow.

>
rate at any point in phase space, rather it moves the fixed point itself. In fact, lowering Cε2
much beyond the value adopted here moves the fixed point outside of the allowable range
0 ≤ f ≤ 1, so that the f = 1 fixed point becomes the attractor.
A similar factor explains why we cannot adjust the model to lower the fixed point on
the upper-left, to try to improve the fit. The location of this point is controlled by the
URANS model coefficients Cε1 < , C< , which are not adjustable parameters of the reduced-
ε2
order model, and must be set to match the coefficients settings used in the URANS we are
comparing to.
Another way we can assess the accuracy of the model is to compare the measured growth
rate in the simulations to the model prediction using Equation (36) with f and g computed
from the simulation data. This is shown in Figure 2. If the model was perfect, all the data
would fall on the dashed line. As is clear from the plot, the agreement improves as the
initial value of f increases. For the two trajectories with the lowest initial f, there are two
distinct regions. The first, corresponding to the initial descent in Figure 1, shows a linear
relationship, but with a significantly different slope, whereas once the trajectory enters the
green shaded region in Figure 1, the measured decay rate is about 30% high. This indicates
that the reduced-order model performs worse in the region above the green shaded region,
which is not surprising, as this is where the URANS deviates the most from physically
realistic behaviour.
Overall, the reduced order model Equations (33) gives us a good feel for the dynamics of
the URANS. We can see that the solution can deviate substantially from the correct decay
rate as it evolves. However, there appears to be a preferred trajectory, a roughly straight
path between two of the fixed points (on the upper left and lower right in Figure 1), along
the lower portion of which the decay rate is close to the experimental values. This may give
a false sense that the URANS approach is correct. At least from the reduced-order model,
it appears that the preferred trajectory is not aligned with a contour of constant decay rate,
although the full simulation trajectories are less clear.

4. Forced turbulence
Decaying turbulence is arguably not analogous to problems, such as bluff body wakes,
in which unsteadiness arises spontaneously in URANS. It is more similar to problems
JOURNAL OF TURBULENCE 383

where the initial condition includes a resolved perturbation or geometric feature that cre-
ates an initial large-scale structure which decays over long times, such as [48,49]. For the
wake, where the flow is globally unstable, the instability acts as a kind of forcing term
which sustains the unsteadiness in the URANS solution. The equivalent mechanism for
homogeneous turbulence would be forced turbulence, which we consider in this section.
Consider a DNS of homogeneous isotropic turbulence generated by a large-scale forc-
ing. The flow evolves under a statistically stationary stochastic forcing until statistical
equilibrium is reached. At this point, as for the decaying problem, the flow is decomposed
into resolved and subfilter quantities using a specified cutoff length scale. The problem
is then restarted, using the URANS equations, initialised with the resolved and modelled
fields, and the evolution is continued with the same forcing. This is again repeated with a
series of different cutoffs.
To model this problem, we will make the following assumptions. First, that the forcing
is entirely in the resolved component, regardless of the cutoff length scale. Second, that
the flow remains in equilibrium, so that we can assume that the forcing is always balanced
by the dissipation, F = ε. Finally, that the forcing does not directly affect the dissipation
or enstrophy. In that case, Equations (27b), (27c) and (27d) remain unchanged, however,
Equation (27a) becomes

∂ k> 
= − P<  − ε>  + F ,
∂t
and
dk
= 0.
dt
The reduced-order model for the forced case is, therefore
df  

=f g−1 , (38)
dt
and the g equation is still Equation (31b). The fixed points are

f =0 g=0 (39a)
C< − 1
f = 0 g = ε2< −1 (39b)
Cε1
 < <

2 Cε1 − Cε2
f = < − 2C< − C> g = 1. (39c)
2Cε1 ε2 ε2

With the standard k − ε coefficients, this results in an attracting fixed point at g = 1 and
f ≈ 0.3, as can be seen in Figure 3.
Unlike the unforced case, the energy partition does not trend towards a steady RANS
solution in the forced case. Instead, the forcing acts as a continuous source of resolved scale
energy. The fixed point represents the balance between the dissipation of resolved scale
motion due to the model and the production of resolved scale motion due to the forcing.
A value of g = 1 at the fixed point means that the energy input due to forcing is the same
as the transfer (production) between resolved and subfilter scales, which is the same as
384 D. M. ISRAEL

Figure 3. Phase space evolution for homogeneous forced turbulence (Equations 38 and 31b). The grey
trajectories are the reduced order model, with the fixed points in black.

the subfilter dissipation, that is, a classic cascade process. The fixed points we obtain are
different from those in [35] because we do not assume that the total dissipation rate follows
the steady RANS equation, as we describe above.

5. Implications for RANS-based hybrid models


The fundamental problem with the URANS approach is that, regardless of how the sim-
ulation is initialised, conventional RANS model equations are designed to account for
all the turbulent scales of motion. When used to simulate only a range of scales, the
dynamics of the solution fundamentally disagrees with the modelling assumptions. To
fix this problem, there exists a class of hybrid turbulence models which, through various
mathematical procedures, attempt to rescale the subfilter quantities in the RANS equa-
tions to properly account for the partition between resolved and modelled scales. These
include the flow-simulation methodology (FSM) [32,34], partially averaged Navier–Stokes
(PANS) [35,36], and the partially integrated transport model (PITM) [39]. The reduced-
order model approach from the previous section is easily extended to these hybrid models.
In fact, for models where the only change is a modification to the model coefficients, the
reduced-order model (31a) still applies, it only needs to be replotted with the appropriate
coefficient values.
To demonstrate this, we consider the PANS model. In the PANS approach, the Cε2 <

coefficient is replaced with a rescaled version,


 < <
 <
Cε2 = fk Cε2 − Cε1 + Cε1 , (40)

fk is a model parameter that is supposed to set the ratio of the modelled to total energy.
This is distinguished from f, which is the actual ratio achieved by the simulation. In other
words, fk is an input to the model and f is a diagnostic; if the PANS model worked exactly
as intended, fk and f would be equal.
Figure 4 shows the trajectory maps for the PANS model with four different values of fk .
The fixed points move depending on fk , and the behaviour is as desired: as fk decreases,
the value of f at the attracting fixed point also decreases. We can substitute Cε2 ∗ from
<
Equation (40) for Cε2 in the expression for the reduced-order model fixed point value of f,
JOURNAL OF TURBULENCE 385

Figure 4. Phase space evolution of unforced PANS with different values of fk . From top left, fk =
0.2, 0.4, 0.6, 0.8.

Figure 5. Phase space evolution for forced PANS with different values of fk . From top left, fk =
0.2, 0.4, 0.6, 0.8.

Equation (34d), to obtain an analytic relationship between fk and f at the fixed point,
 < <

2 Cε1 − Cε2
f = < − C> − 2 fk . (41)
2Cε1 ε2

For the standard values of the model coefficients, this is approximately f ≈ 0.96fk , which
is very close to the desired equality. Note also, that from Equation (37), the growth rate at
the fixed point is independent of fk .
We can conduct the same experiment for the forced case. In this case, we do not expect
the result to relax to a steady solution in the limit of fk → 1, since the underlying RANS
model does not do so. Instead, the level of unsteadiness is restricted to range of f ≈ 0 − 0.3,
as can be seen in Figure 5.
In other words, for the unforced case, the PANS model does exactly what it is designed
to do. Unlike URANS, where the limiting behaviour relaxes to a conventional RANS, the
386 D. M. ISRAEL

PANS model tends towards a value of f very close to the value of the input parameter fk .
Nevertheless, it is important to be aware that the PANS model also can have significant
deviations from the fixed point, particularly if the initial condition is not chosen so that
both f and g are at the fixed point value for the specified fk . For the forced case, the rela-
tionship between the input fk and the observed f is more complex. Obviously, the PANS
model cannot enforce a steady result at the fk = 1 limit if the underlying RANS model
cannot.

6. Conclusion
Conventional RANS models were developed using a single-scale assumption. Such closures
are only justified where any resolved unsteadiness exhibits a ‘spectral gap ’, or, equivalently,
the large scales, such as the total turbulent kinetic energy and the integral length scale,
completely characterise the turbulence. Even for problems with large unsteady turbulent
structures, these models were designed to produce the correct solution for the time mean,
when solved for the steady-state solution (i.e. when the time-derivative terms are set to
zero). Proponents of the URANS approach have not produced a rigorous justification as to
why these models sometimes give rise to unsteady solutions if allowed to evolve in time,
nor what precisely those unsteady solutions represent. Without clearly identifying what
the models should do, it is hard to say definitely whether the results they do produce are
correct or not, or even what would be the correct way to employ them.
One key limitation in our understanding is that existing research into URANS consists
of case studies. Even very careful analysis can lead to uncertain, and even contradictory,
conclusions, given the enormous parameter space of all potential applications of URANS.
What has been missing is a theoretical framework for understanding what URANS actually
does. This paper presents a first step in that direction.
The proposed model can be viewed as an extension of earlier fixed point analyses. The
power of the approach lies in viewing the evolution of the URANS simulation as a dynam-
ical system, with the phase space dynamics giving insight into the global behaviour of the
model. The purpose of the model is not to be an accurate predictive tool, but rather to
reproduce the qualitative behaviours we expect to see in URANS simulations. Comparisons
to actual simulation data indicate that the model does just that.
Without a clear metric for correctness, it is impossible to say definitively that URANS
is wrong, but these results show that the idea that a steady RANS model can simply be
applied to unsteadiness and a correct answer obtained is a myth. Defenders of the URANS
approach might try to argue that the results in Figure 1, showing that the URANS has
a preference for trajectories that fall near the region of correct decay rate, vindicates the
approach. However, there are three serious difficulties with this view. First, the fact that
a URANS solution will decay in time until almost all resolved scales are dissipated and a
steady RANS solution is achieved, is neither a useful model behaviour, nor consistent with
most users expectations for an unsteady model. Second, both the stick spectrum analysis
and the simulation data initialised with forced turbulence show that initialising the URANS
so that it starts in a state where the decay rate is reasonable is not a trivial undertaking.
Finally, while the model may, in certain cases, predict the decay rate properly, in others it
can undergo extreme transients into highly non-physical behaviours.
Extending the analysis to the hybrid PANS model, we see that for decaying turbulence
PANS solves the first of these issues, placing the fixed point at exactly the resolution the
JOURNAL OF TURBULENCE 387

user requested. However, the path to the fixed point, and, in particular, inconsistent initial
conditions, still are issues that require further study. Here too, the forced case shows more
complicated dynamics. This is consistent with the observation that the hybrid model can-
not be calibrated to enforce an energy partition more coarse than the underlying RANS
model upon which it is built.
The approach does need to be extended to more common applications of URANS. Work
is underway for an extension to free shear and mixing layers. In the meantime, the case of
forced turbulence provides a possible analog for the mechanisms at work in URANS of
globally unstable flows, with the forcing acting as a proxy for the global instability. These
results show that in such a case the forcing can keep the unsteadiness alive. The energy
partition is controlled by a balance between the model damping and the forcing input.
In either case, further analysis would benefit if the proponents of the URANS approach
clearly defined metrics for what they believe the correct behaviour of the model should
be. In the meantime, users of models would be advised to have a very clear understanding
of what can and cannot be expected of RANS models before applying them to unsteady
problems.

Notes
1. We can slightly generalise this to include any parameter that defines the decomposition into
resolved and modelled energy. The PANS approach [36], for example, parameterises the decom-
position by the fraction of energy which is modelled, rather than the cutoff length scale.
Similarly, simulation approaches that use explicit filtering [50,51], which are parameterised by
a filter width, and implicit LES (ILES), for which the numerical viscosity which is used as the
turbulence model is a function of the grid spacing, all fall within the Travin et al. [31] definition
of LES.
2. This is consistent with the usage of Pope [8], but contrary to Spalart [30], who takes URANS
and VLES as synonymous.

Acknowledgments
The author would like to express his gratitude to Dr. Colin Towery for providing simulation data for
comparison, as well as very helpful discussions. Thanks also to Dr. Filipe Pereira for his insightful
comments on the manuscript. The reviewers provided very helpful feedback that greatly improved
the presentation. This research was funded by the LANL Mix and Burn project under the DOE ASC,
Physics and Engineering Models program.

Disclosure statement
No potential conflict of interest was reported by the author(s).

ORCID
Daniel M. Israel http://orcid.org/0000-0001-8162-1885

References
[1] Reynolds O. On the dynamical theory of incompressible viscous fluids and the determination
of the criterion. Philos Trans R Soc Lond A. 1895;186:123–164. doi:10.1098/rsta.1895.0004
[2] Taylor GI. Eddy motion in the atmosphere. Philos Trans R Soc Lond A. 1915;215:1–26.
doi:10.1098/rsta.1915.0001
388 D. M. ISRAEL

[3] Prandtl L. Bericht über untersuchungen zur ausgebildeten turbulenz. ZAMM J Appl Math
Mech/Z Angew Math Mech. 1925;5(2):136–139. doi: 10.1002/zamm.19250050212
[4] Smagorinsky J. General circulation experiments with the primitive equations: I. The basic
experiment. Mon Weather Rev. 1963 Mar;91(3):99–164. doi: 10.1175/1520-0493(1963)091 < 00
99:GCEWTP > 2.3.CO;2
[5] Lilly DK. The representation of small-scale turbulence in numerical simulation experiments.
In: Proceedings of the IBM Scientific Computing Symposium on Environmental Sciences;
Yorktown Heights, IBM, NY; 1967. p. 195–210.
[6] Deardorff JW. A numerical study of three-dimensional turbulent channel flow at large Reynolds
numbers. J Fluid Mech. 1970;41(part 2):453–480. doi: 10.1017/S0022112070000691
[7] Germano M. Turbulence: the filtering approach. J Fluid Mech. 1992;238:325–336. doi:
10.1017/S0022112092001733
[8] Pope SB. Turbulent flows. Cambridge: Cambridge University Press; 2000.
[9] Celenligil MC, Mellor GL. Numerical solution of two-dimensional turbulent separated
flows using a Reynolds stress closure model. J Fluids Eng. 1985 Dec;107(4):467–476. doi:
10.1115/1.3242515
[10] Lasher WC, Taulbee DB. On the computation of turbulent backstep flow. Int J Heat Fluid Flow.
1992;13(1):30–40. doi: 10.1016/0142-727X(92)90057-G
[11] Claus RW, Huang PG, MacInnes JM. Time-accurate simulations of a shear layer forced at a
single frequency. AIAA J. 1990;28(2):267–275. doi: 10.2514/3.10384
[12] Launder BE, Morse A, Rodi W, et al. Prediction of free shear flows: a comparison of the per-
formance of six turbulence models. In: Free Turbulent Shear Flows: Volume I-Conference
Proceedings; NASA; 1973. p. 361–426; SP-321. Proceedings of a conference held at NASA
Langley Research Center, Hampton, Virginia, Jul 20–21, 1972.
[13] Launder BE, Spalding DB. The numerical computation of turbulent flows. Comput Methods
Appl Mech. 1974;3(2):269–289. doi: 10.1016/0045-7825(74)90029-2
[14] Franke R, Rodi W. Calculation of vortex shedding past a square cylinder with various turbu-
lence models. In: Durst F, Friedrich R, Launder BE, et al., editors. Turbulent Shear Flows 8;
Berlin, Heidelberg. Springer Berlin Heidelberg; 1993. p. 189–204.
[15] Rodi W. On the simulation of turbulent flow past bluff bodies. In: Murakami S, editor. Compu-
tational Wind Engineering 1. Oxford: Elsevier; 1993. p. 3–19. https://www.sciencedirect.com/
science/article/pii/B978044481688750005X.
[16] Bosch G, Rodi W. Simulation of vortex shedding past a square cylinder with different turbu-
lence models. Int J Numer Methods Fluids. 1998;28(4):601–616. doi: 10.1002/(ISSN)1097-0363
[17] Launder BE, Reece GJ, Rodi W. Progress in the development of a Reynolds-stress turbulence
closure. J Fluid Mech. 1975;68(3):537–566. doi: 10.1017/S0022112075001814
[18] Johansson SH, Davidson L, Olsson E. Numerical simulation of vortex shedding past triangular
cylinders at high Reynolds number using a k − ε turbulence model. Int J Numer Methods
Fluids. 1993;16(10):859–878. doi: 10.1002/fld.1650161002
[19] Jansson LS, Davidson L, Olsson E. Calculation of steady and unsteady flows in a film-cooling
arrangement using a two-layer algebraic stress model. Num Heat Trans A. 1994;25(3):237–258.
doi: 10.1080/10407789408955947
[20] Durbin PA. Separated flow computations with the k − ε − v2 model. AIAA J. 1995;33(4):
659–664. doi: 10.2514/3.12628
[21] Travin A, Shur M, Strelets M, et al. Detached-eddy simulations past a circular cylinder. Flow
Turb Comb. 2000;63(1):293–313. doi: 10.1023/A:1009901401183
[22] Iaccarino G, Ooi A, Durbin P, et al. Reynolds averaged simulation of unsteady separated flow.
Int J Heat Fluid Flow. 2003;24(2):147–156. doi: 10.1016/S0142-727X(02)00210-2
[23] Squires KD, Forsythe JR, Spalart PR. Detached-eddy simulation of the separated flow over a
rounded-corner square. J Fluids Eng. 2005 Sep;127(5):959–966. doi: 10.1115/1.1990202
[24] Shur M, Spalart PR, Squires KD, et al. Three-dimensionality in Reynolds-averaged
Navier–Stokes solutions around two-dimensional geometries. AIAA J. 2005;43(6):1230–1242.
doi: 10.2514/1.9694
[25] Jin G, Braza M. Two-equation turbulence model for unsteady separated flows around airfoils.
AIAA J. 1994;32(11):2316–2320. doi: 10.2514/3.12292
JOURNAL OF TURBULENCE 389

[26] Carpy S, Manceau R. Turbulence modelling of statistically periodic flows: Synthetic jet into
quiescent air. Int J Heat Fluid Flow. 2006;27(5):756–767. Special issue of the 6th Interna-
tional Symposium on Engineering Turbulence Modelling and Measurements – ETMM6.
10.1016/j.ijheatfluidflow.2006.04.002
[27] Fadai-Ghotbi A, Manceau R, Borée J. Revisiting URANS computations of the backward-
facing step flow using second moment closures. Influence of the numerics. Flow Turb Comb.
2008;81(3):395–414. doi: 10.1007/s10494-008-9140-8
[28] Palkin E, Mullyadzhanov R, Hadžiabdić M, et al. Scrutinizing URANS in shedding
flows: the case of cylinder in cross-flow in the subcritical regime. Flow Turb Comb.
2016;97(4):1017–1046. doi: 10.1007/s10494-016-9772-z
[29] Pereira FS, Eça L, Vaz G, et al. On the simulation of the flow around a circular cylinder at
Re = 140, 000. Int J Heat Fluid Flow. 2019;76:40–56. doi: 10.1016/j.ijheatfluidflow.2019.01.007
[30] Spalart P. Strategies for turbulence modelling and simulations. Int J Heat Fluid Flow.
2000;21(3):252–263. doi: 10.1016/S0142-727X(00)00007-2
[31] Travin A, Shur M, Spalart P, et al. On URANS solutions with LES-like behaviour. In: Neit-
taanmäki P, Rossi T, Majava K, et al., editors. Congress on Computational Methods in Applied
Sciences and Engineering ECCOMAS; 2004.
[32] Speziale CG. Turbulence modeling for time-dependent RANS and VLES: a review. AIAA J.
1998 Feb;36(2):173–184. doi: 10.2514/2.7499
[33] Israel DM, Postl D, Fasel HF. A flow simulation methodology for simulation of coherent
structures and flow control. AIAA; 2004. AIAA Paper 2004–2225. 2nd AIAA Flow Control
Conference, Jun 28–Jul 1, 2004, Portland, OR.
[34] Fasel HF, von Terzi DA, Sandberg RD. A methodology for simulating compressible turbulent
flows. J Appl Mech. 2006 May;73(3):405–412. doi: 10.1115/1.2150231
[35] Girimaji SS, Jeong E, Srinivasan R. Partially averaged Navier–Stokes method for turbulence:
fixed point analysis and comparison with unsteady partially averaged Navier–Stokes. J Appl
Mech. 2006 May;73(3):422–429. doi:10.1115/1.2173677
[36] Girimaji SS. Partially-averaged Navier–Stokes model for turbulence: a Reynolds-averaged
Navier–Stokes to direct numerical simulation bridging method. J Appl Mech. 2006
May;73:413–421. doi: 10.1115/1.2151207
[37] Girimaji SS, Abdol-Hamid KS. Partially-averaged Navier Stokes model for turbulence: imple-
mentation and validation; 2005. AIAA Paper 2005–502. 43rd AIAA Aerospace Sciences
Meeting & Exhibit, Jan 10–13, 2005, Reno, NV.
[38] Chaouat B, Schiestel R. A new partially integrated transport model for subgrid-scale stresses
and dissipation rate for turbulent developing flows. Phys Fluids. 2005;17(6):065106. doi:
10.1063/1.1928607
[39] Schiestel R, Dejoan A. Towards a new partially integrated transport model for coarse grid and
unsteady turbulent flow simulations. Theor Comput Fluid Dyn. 2005 Feb;18(6):443–468. doi:
10.1007/s00162-004-0155-z
[40] Schumann U. Subgrid scale model for finite difference simulations of turbulent flows in plane
channels and annuli. J Comp Phys. 1975 Aug;18(4):376–404. doi: 10.1016/0021-9991(75)900
93-5
[41] Yoshizawa A. A statistically-derived subgrid model for the large-eddy simulation of turbulence.
Phys Fluids. 1982 Sep;25(9):1532–1538. doi: 10.1063/1.863940
[42] Mohamed MS, Larue JC. The decay power law in grid-generated turbulence. J Fluid Mech.
1990;219:195–214. doi: 10.1017/S0022112090002919
[43] Lavoie P, Djenidi L, Antonia RA. Effects of initial conditions in decaying turbulence generated
by passive grids. J Fluid Mech. 2007;585:395–420. doi: 10.1017/S0022112007006763
[44] Speziale CG, Mhuiris NMG. On the prediction of equilibrium states in homogeneous turbu-
lence. J Fluid Mech. 1989;209:591–615. doi:10.1017/S002211208900323X
[45] Haering SW, Oliver TA, Moser RD. Active model split hybrid RANS/LES. Phys Rev Fld. 2022
Jan;7:014603. doi: 10.1103/PhysRevFluids.7.014603
[46] Perot JB, Kops SMDB. Modeling turbulent dissipation at low and moderate Reynolds numbers.
J Turb. 2006;7:N69. doi: 10.1080/14685240600907310
390 D. M. ISRAEL

[47] Towery CAZ, Sáenz JA, Livescu D. Posterior comparison of energy partitioning dynamics in
several hybrid RANS-LES turbulence models. In preparation. 2023.
[48] Morgan BE, Greenough JA. Large-eddy and unsteady RANS simulations of a shock-accelerated
heavy gas cylinder. Shock Waves. 2016;26(4):355–383. doi: 10.1007/s00193-015-0566-3
[49] Haines BM, Grinstein FF, Schwarzkopf JD. Reynolds-averaged Navier–Stokes initializa-
tion and benchmarking in shock-driven turbulent mixing. J Turb. 2013;14(2):46–70. doi:
10.1080/14685248.2013.779380
[50] Bogey C, Bailly C. Effects of inflow conditions and forcing on subsonic jet flows and noise.
AIAA J. 2005;43(5):1000–1007. doi: 10.2514/1.7465
[51] Mathew J, Lechner R, Foysi H, et al. An explicit filtering method for large eddy simulation of
compressible flows. Phys Fluids. 2003;15(8):2279–2289. doi: 10.1063/1.1586271
[52] Corrsin S. The isotropic turbulent mixer: part II. arbitrary Schmidt number. AIChE J.
1964;10(6):870–877. doi: 10.1002/aic.690100618
[53] Comte-Bellot G, Corrsin S. The use of a contraction to improve the isotropy of grid-generated
turbulence. J Fluid Mech. 1966;25(4):657–682. doi: 10.1017/S0022112066000338
[54] Skrbek L, Stalp SR. On the decay of homogeneous isotropic turbulence. Phys Fluids.
2000;12(8):1997–2019. doi: 10.1063/1.870447
[55] Ristorcelli JR. Passive scalar mixing: analytic study of time scale ratio, variance, and mix rate.
Phys Fluids. 2006 Jul;18(7):075101. doi: 10.1063/1.2214704

Appendices
Appendix 1. Stick spectrum analysis
The initial conditions for a URANS-type simulation (or, equivalently, for a PANS/hybrid/SRS simu-
lation) cannot fall in any arbitrary point in f, g, R space. The production-to-dissipation ratio, g, will
depend on the energy partition, f. Another way to look at it is, given a particular filter that is param-
eterised by a length scale, , for each we will have specific values of f, g, and R. That is, f = f ( ),
g = g( ), and R = R( ) are all functions of . Choosing a different filter shape will result in a slightly
different curve, but it is likely that this will be a weak effect.
We can estimate this dependence by using an assumed spectrum analysis [52–55]. If we take L as
the characteristic length scale of the large scales, and the Kolmogorov scale, η, as characterising the
smallest scales, we can use an assumed spectrum of

C ε2/3 κ −5/3 κ ∈ [2π/L, 2π/η]
E (κ) = . (A1)
0, elsewhere

The subfilter turbulent kinetic energy is


 ∞
k<  = E (κ) dκ. (A2)

Integrating the assumed spectra Equation (A1) gives an estimate of


 2/3   η 2/3 
3
k<  = C ε2/3 1− . (A3)
2 2π

The total turbulent kinetic energy is just the RANS limit of Equation (A3): when goes to zero
(actually = L) all scales are captured, so
 2/3   η 2/3 
3 L
k = C ε2/3 1− . (A4)
2 2π L
JOURNAL OF TURBULENCE 391

Using this we can obtain an estimate for f,


 
 2/3 1 −  η 2/3
f =   2/3  . (A5)
L 1 − ηL

To compute g and R, we need the subfilter enstrophy,


 ∞
 2
ω = κ 2 E (κ) dκ. (A6)

We can use our assumed spectrum to obtain


 4/3   η 4/3 
 3 2π
ω2 = C ε2/3 1− , (A7)
4 η
and, therefore,
  η 2/3 2
 4/3 
ε2/3 η 4/3 1 −
R = 3Cμ C 2   4/3  . (A8)
ν 2π 2π 1− η

To find the leading order scaling of R, we can eliminate ε and ν in terms of L and η. To do this, we
cannot use the usual dimensional estimates, which are not correct for our stick spectrum. Instead,
we must use Equation (A4) and
 4/3   η 4/3 
 2  2 3 2/3 2π
ε = ν ω = ν lim ω< = Cν ε 1− . (A9)
→∞ 4 η L
With these we find that R scales with Ret ,
  2
 4/3 1 −  η 4/3 1 −  η 2/3
L
R = Cμ Ret   η 4/3    2/3 2 . (A10)
L 1− 1− η L

Finally,
 <  2  2
P<  ν ω̄ ω̄
g= ≈ T 2 = R  2 , (A11)
ε<  ν ω< ω<
or
   2/3 2
 4/3    η 4/3 2 1 − η
27
g = Cμ C3 1 − 1−   4/3 2 .
16 L L
1− η

If we further consider the case when η → 0, then


 2/3
f =
L
  4/3 
27 3
g = Cμ C 1 − ,
16 L
or
27  
g= Cμ C3 1 − f 2 . (A12)
16
392 D. M. ISRAEL

Appendix 2. URANS Simulations


The URANS simulations used to validate the reduced-order model were performed by colleagues at
Los Alamos National Laboratory as part of a wider study of hybrid RANS-LES methods; a complete
report of their work is forthcoming [47]. A brief description of the relevant details of simulations
follows. All the simulations were performed using an incompressible, pseudo-spectral code with a
triply-periodic domain of size 2π × 2π × 2π. The code uses the two-thirds rule for dealiasing and
a fourth-order explicit Runge–Kutta with a CFL based dynamic time step for time advancement.
The DNS data for homogeneous isotropic decaying turbulence used to generate the initial con-
ditions for the URANS were created starting with a 7683 physical space mesh with random forcing
in a range of wave numbers, 4 ≤ κ ≤ 6. After this simulation reached a statistically steady state, the
simulation was restarted on a 15633 domain and ran for two integral time-scales, and then again on
a 20483 domain for an additional integral time scale. At this point, the forcing was turned off and
the turbulence was allowed to decay for two integral time-scales, to achieve a power-law decay.
To obtain the initial conditions for the URANS, a Gaussian filter was applied, and both the fil-
tered velocity fields and k< and ε< were computed. The RANS equations were solved in logarithmic
formulations (solving for ln k< and ln ε< ) to ensure positivity and improve stability. The DNS fields
used to generate the initial conditions were resolved to half the Kolmogorov scale, and so are very
well resolved. The URANS simulations were performed with a grid resolution of one-third the filter
width used to generate the initial conditions.

You might also like