You are on page 1of 29

PHYSICS OF FLUIDS VOLUME 16, NUMBER 12 DECEMBER 2004

The effect of subgrid-scale models on the vortices computed


from large-eddy simulations
Carlos B. da Silvaa) and José C. F. Pereira
Instituto Superior Técnico, Pav. Mecânica I, 1° andar/esq. (LASEF), Avenida Rovisco Pais,
1049-001 Lisboa, Portugal
(Received 19 December 2003; accepted 2 September 2004; published online 4 November 2004)
Direct numerical and large-eddy simulations (DNS/LES) of temporal plane jets are carried out in
order to analyze the effect of the subgrid-scale (SGS) models on the vortices obtained from LES.
The dynamics of the filtered vorticity norm (or filtered enstrophy) is analyzed through the
application of a box filter to temporal DNS of turbulent plane jets 共Re␭ ⬇ 100兲, using a methodology
similar to da Silva and Métais [J. Fluid Mech. 473, 103 (2002)]. Special emphasis is placed on the
enstrophy SGS dissipation term, which represents the effect of the SGS models on the vortices
computed from LES. When the filter is placed in the inertial range region the evolution of the
vorticity norm is governed by the enstrophy production and enstrophy SGS dissipation, which
represents, in the mean, a sink of resolved enstrophy. Thus the coherent vortices obtained from LES
are subjected to an additional (nonviscous) dissipation mechanism. Locally, however, the enstrophy
SGS dissipation can be either a sink or a source of resolved vorticity (forward/backward enstrophy
cascade), but the forward cascade dominates, in analogy with what happens with the resolved
kinetic energy equation. A priori tests are conducted using several SGS models in order to analyze
their ability to represent the enstrophy SGS dissipation. The models analyzed are the Smagorinsky,
structure function, filtered structure function, dynamic Smagorinsky, gradient, scale similarity, and
mixed. It turns out that in terms of spatial localization all the models lead to a good correlation
between the “real” and “modeled” enstrophy SGS dissipation. Moreover, all the SGS models, even
of eddy-viscosity type, are able to provide enstrophy SGS backscatter. However, in terms of
statistical behavior the eddy-viscosity models do not provide enough enstrophy backscatter as the
non-eddy-viscosity models. LES are carried out and show that the Smagorinsky, structure function,
and mixed models cause excessive vorticity dissipation compared to the other models, and although
the enstrophy SGS dissipation affects mainly the smallest resolved scales, it may affect also some
low-wave numbers. An estimation of the “vorticity error” and its wave number dependence is given,
for each SGS model. Both a priori tests and LES show that the dynamic Smagorinsky and filtered
structure function models seem to be the best suited to a correct prediction of the resolved vorticity
field. © 2004 American Institute of Physics. [DOI: 10.1063/1.1810524]

I. INTRODUCTION However, the influence of the SGS models on the vorti-


Large-eddy simulation (LES) is now a well established ces computed from LES has been mostly neglected. For in-
tool currently used by researchers in the study of complex stance, few studies analyzed the ability of the SGS models to
turbulent flows (Lesieur and Métais,1 Meneveau and Katz2). reproduce the “correct” size, shape, and lifetime of these
Although there are a great deal of LES studies in which the structures in a given flow. It has been assumed that the large-
statistical characteristics of turbulent flows (in the sense of scale vortices, which are the ones of interest in engineering
Meneveau3) are used to evaluate the subgrid-scale (SGS) applications, are explicitly computed in LES and therefore
model performance, there are comparatively fewer studies the effect of the SGS models on these structures is negli-
which assess the model’s ability to reproduce some relevant gible. However, some researchers have observed that differ-
instantaneous flow features—e.g., the flow structures. Since ent SGS models produce differences in the coherent vortices
the kinetic energy exchange between the resolved and unre- computed from LES (e.g., Silvestrini,7 Vreman, Geurts, and
solved scales is a key feature that the SGS models must Kuerten8), which shows unambiguously that there is an in-
represent, several authors have examined the influence of the fluence of the SGS models on the vortices computed from
coherent vortices on the so-called “subgrid-scale dissipation” LES. To the authors’ knowledge, there exists no systematic
term ␶ijS⬍ 4 5
ij (e.g., Piomelli et al., O’Neil and Meneveau ). In
study on the nature of this mechanism.
this context, a complete account of the role played by the The present work focuses on analyzing this problem
coherent vortices on the interscale interactions in plane jets through a priori and a posteriori tests of temporal evolving
was recently given by da Silva and Métais.6 plane jets. Although some new very promising SGS models/
techniques have emerged recently (e.g., Domaradzki and
a)
Telephone: (⫹351) 21-841-73-68. Fax: (⫹351) 21-849-52-41. Electronic Saiki,9 Domaradzki and Loh,10 Domaradzki, Loh, and Yee,11
mail: csilva@navier.ist.utl.pt Geurts,12 Winckelmans et al.,13 Stolz, Adams, and

1070-6631/2004/16(12)/4506/29/$22.00 4506 © 2004 American Institute of Physics

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4507

Kleiser,14,15 Brun and Friedrich16), we restrict our study to To analyze the coherent vortices present in large-eddy
the more commonly used SGS models, since these are the simulations (LES), we have to consider the filtered Navier–
ones which are more likely to interest present day research- Stokes equations, from which one can derive the resolved
ers. The models analyzed here are the Smagorinsky enstrophy transport equation,
(SMAG), structure function (SF), filtered structure function
(FSF), dynamic Smagorinsky (DSMAG), gradient (GRAD),
scale similarity (SS), and mixed (MIX) models.
We begin this study by analyzing the dynamics of the
resolved vorticity norm (or enstrophy) in a direct numerical
simulation (DNS) of a temporal plane jet. The enstrophy
dissipation caused by the filtering operation applied to the
Navier–Stokes equations is described by an enstrophy SGS
dissipation term. After this, the SGS models are analyzed
using classical a priori tests, but focusing on the enstrophy
related issues. Finally, several LES are carried out and com-
pared with results from the filtered DNS data.
It should be noted that the present work focuses on the 共2兲
vortices from the far field of fully developed turbulent plane
jets. Plane jets were chosen in this study because they exhibit
well documented vortices which are similar to the ones Here “⬍” denotes a (spatial) filtering operation defined by
found in other free shear layers (e.g., mixing layers and
wakes), provided they are at the far field region. Therefore,


the conclusions from the present work can be extended to
vortices from the far field of free shear layers. ␾⬍共x兲 = ␾共x⬘兲G⌬共x − x⬘兲dx⬘ , 共3兲
The paper is organized as follows. In Sec. II the main ⍀
equations for the transport of the grid-scale enstrophy are
derived and the SGS models analyzed in this study are re-
viewed. Section III describes the numerical code and lists the where ␾ is a given flow variable and G⌬共x兲 is the filter ker-
parameters for all the simulations carried out in this study. nel. The filtering operation yields the subgrid-stress tensor,
The DNS of a temporal plane jet is validated in Sec. IV. The ␶ij = 共uiu j兲⬍ − u⬍ ⬍
i uj .
physical mechanism of the enstrophy SGS dissipation is ana- In Eq. (2) the terms I and II account for the total (local
lyzed in Sec. V through the application of a box filter to the and convective) variation of resolved enstrophy, due to the
plane jet DNS. After that (Sec. VI) the modeled enstrophy resolved enstrophy production (term III), resolved viscous
SGS dissipation from each SGS model is analyzed and com- diffusion (IV), resolved viscous dissipation (V), and to an
pared with the one from the filtered DNS. Finally (Sec. VII), additional term: the enstrophy SGS dissipation (VI).
several LES are carried out with all the SGS models, in order This term originates in the filtering operation applied to
to validate a posteriori the previous analysis. Section VIII the Navier–Stokes equations and thus represents a transfer of
closes the work with an overview of its main results and enstrophy between grid (or resolved) and subgrid-scale en-
conclusions. strophy. If ␧ijk⍀⬍ i ⳵ ␶kp / ⳵x j ⳵ x p ⬍ 0 the term represents a loss
2

of enstrophy (forward enstrophy cascade) from grid scales


into subgrid scales, whereas ␧ijk⍀⬍ i ⳵ ␶kp / ⳵x j ⳵ x p ⬎ 0 implies
2

II. GOVERNING EQUATIONS an inverse transfer (backward enstrophy cascade) of enstro-


A. Transport equation for the resolved enstrophy phy from the subgrid into the resolved scales of motion.
Thus, the influence of the SGS models in the vortical struc-
In turbulent flows, a coherent vortex implies, by defini- tures computed from LES is described by this term, which is
tion, high local vorticity.53 Therefore, the evolution of these zero in DNS where Eq. (2) reverts to Eq. (1).
structures is governed by the enstrophy transport equation For subsequent analysis, it is interesting to split the en-
(Chassaing17), strophy SGS dissipation (VI) into a source (enstrophy SGS
D 1
冉 冊
Dt 2
⍀i⍀i = ⍀i⍀ jSij + ␯
⳵2 1
⳵ xj ⳵ xj 2
冉 冊
⍀ i⍀ i
transfer—VII) and diffusion (enstrophy SGS diffusion—
VIII) terms,18

⳵ ⍀i ⳵ ⍀i
−␯ , 共1兲
⳵ xj ⳵ xj
where ui and ⍀i are the velocity and vorticity vectors, re-
spectively, Sij = 21 共⳵ui / ⳵x j + ⳵u j / ⳵xi兲 is the rate of strain tensor
and ␯ is the molecular viscosity. Equation (1) stems directly
from the Navier–Stokes equations and is valid both locally
and instantaneously. 共4兲

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4508 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

B. Subgrid-scale models F2共xជ ,⌬x兲 = 具储uជ 共xជ + rជ兲 − uជ 共xជ 兲储典储r储=⌬x

冕 冋 册
1. Smagorinsky model ⬁
sin k⌬x
19
The Smagorinsky model (SMAG) is the most famous =4 E共k兲 1 − dk, 共9兲
0 k⌬x
and still the most widely used model. It is based on an eddy-
viscosity assumption, which relates the three-dimensional kinetic energy spectrum
E共k兲 to the second-order structure function F2共xជ , ⌬x兲 in ho-
␶ij − 31 ␦ij␶kk = − 2␯tS⬍
ij , 共5兲 mogeneous isotropic turbulence. Before putting the spectral
turbulent eddy viscosity into the physical space it is conve-
and in the local equilibrium hypothesis of the subgrid scales. nient to remove its wave number dependence. Discarding the
Using the isotropic turbulence estimate, cusp behavior (Lesieur and Cholet25) and assuming a balance
3 between the total SGS viscous dissipation and the total (iso-
qsgs
␧⬃ 共6兲 tropic) kinetic energy dissipation (Leslie and Quarini26) one
l gets
(in which the viscous dissipation ␧ is related to a local ve-
locity qsgs and length scale l of the turbulence) and Eq. (5),
the turbulent viscosity in the Smagorinsky model is given by
2
␯t共k兩kc,t兲 = CK−3/2
3
E共kc,t兲
kc
冋 册 1/2
, 共10兲

where E共kc , t兲 is the kinetic energy spectrum E共k兲 at the cut-


␯t共xជ ,t兲 = 共Cs⌬兲2兩S⬍兩, 共7兲 off wave number kc. The most difficult part remains the
transposition of this spectral turbulent eddy viscosity into the
where the velocity and length scales of turbulence were ap- physical space. If one supposes, for each time and at each
proximated by qsgs ⬃ l兩S⬍兩 and l ⬃ ⌬, respectively. Cs is a spatial location, that the turbulence is “locally” homoge-
model constant, ⌬ is a characteristic length scale of the sub- neous, one can relate the “local” three-dimensional kinetic
grid scales and usually taken to be the width of the implicit energy spectrum to a local filtered second-order structure
grid filter ⌬ = 共⌬x ⫻ ⌬y ⫻ ⌬z兲1/3, and 兩S⬍兩 is the norm of the function F⬍ ជ , ⌬x兲 through a modified version of
2 共x
rate of deformation tensor 兩S⬍兩 = 共2S⬍ ⬍ 1/2
ij Sij 兲 . 24
Batchelor’s formula,
In isotropic turbulence, the constant Cs can be related to
the Kolmogorov constant CK through F⬍ ជ ,t,⌬x兲 = 具储uជ ⬍共xជ + rជ,t兲 − uជ ⬍共xជ ,t兲储典储r储=⌬
2 共x

Cs ⬇ 冉 冊
1 2
␲ 3CK
3/4
= 0.18. 共8兲 =4 冕 kc

0

E共k,t兲 1 −
sin k⌬x
k⌬x
dk.册 共11兲

In practice this constant is flow dependent. Here we used The assumption of a Kolmogorov spectrum in Eqs. (10)
Cs = 0.13, since it seems to be the best value to use in free and (11) leads to the SF model,
shear layers (Vreman, Geurts, and Kuerten8). ␯SF ជ ,t兲 = 0.105CK−3/2⌬x关F⬍ ជ ,⌬兲兴1/2 ,
t 共x 2 共x 共12兲
The drawbacks of this model, which have been widely
studied, are that the model is usually too dissipative, ob- in which CK is the Kolmogorov constant, here taken to be
serves a poor correlation with the actual stresses, and is al- 1.4, and F⬍ ជ , ⌬x兲 is a local structure function of the filtered
2 共x
ways a purely dissipative model, i.e., does not allow kinetic velocity field calculated as
energy backscatter (Meneveau and Katz,3 Vreman, Geurts,
and Kuerten,8 O’Neil and Meneveau6). Furthermore, in com- F⬍ ជ ,⌬x兲 = 具储uជ ⬍共xជ + rជ,t兲 − uជ ⬍共xជ ,t兲储典储r储=⌬x .
2 共x 共13兲
plex flows such as in rotating frames or near separating re- This model works well in isotropic turbulence (Métais
gions, the global equilibrium assumption of the subgrid and Lesieur23), but is too dissipative in inhomogeneous flows
scales is not verified (the local equilibrium assumption fails due to its strong dependence from the large-scale inhomoge-
even in simple flow configurations as shown by da Silva and neities. To deal with this problem several variants of the
Métais1). However and surprisingly, the Smagorinsky model structure function model have been developed.
predicts very well the global kinetic energy dissipation of the
energy cascade found in turbulent shear flows (Shao, Sarkar,
and Pantano20).
3. Filtered structure function model
The FSF model is defined by applying a Laplacian filter
2. Structure function model to the velocity field (to remove the contribution of large-
scale inhomogeneities) before calculating the local second-
The main idea of the structure function model consists in order structure function. The final formulation is (Ducros,
transposing the spectral eddy viscosity, in the frame of the Comte, and Lesieur27),
isotropic EDQNM analytical theory of turbulence
(Kraichnan,21 Lesieur22), into the physical space. This was ␯FSF ជ ,t兲 = 0.0014CK−3/2⌬x关F̃⬍ ជ ,⌬x兲兴1/2 ,
t 共x 2 共x 共14兲
made by Métais and Lesieur.23 The starting point is the rela-
tion by Batchelor,24 where

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4509

F̃⬍ ជ ,⌬x兲 = 具储L3共uជ ⬍共xជ + rជ,t兲兲 − L3共uជ ⬍共xជ ,t兲兲储典储r储=⌬x ally combined with the Smagorinsky model, becoming what
2 共x 共15兲
is called the mixed model.
and the Laplacian filter (iterated three times) is defined as
L共uជ ⬍兲 = 兺i 关uជ ⬍共xជ − rជ,t兲 − 2uជ ⬍共xជ,t兲 + uជ ⬍共xជ + rជ,t兲兴. 共16兲
6. Gradient model
Whereas the structure function model represents a slight The gradient or nonlinear model (GRAD) expresses ␶ij
improvement over the Smagorinsky model, the filtered struc- as an inner product of the velocity gradient tensors,
ture function model combines the advantages of low compu- ⬍
⳵ u⬍
i ⳵ uj
tational cost without the problems of excessive dissipation ␶ij = cA⌬2 , 共21兲
thus making it possible to study transitional flows (e.g., ⳵ xk ⳵ xk
da Silva and Métais28). with cA = 1 / 12. The gradient model is equivalent to the scale-
similarity model with ␣ = 1 (de Borue and Orszag32), but it is
4. Dynamic Smagorinsky model much more efficient in the simulations since no extra filter-
ing is needed.
In the dynamic Smagorinsky model (DSMAG) a dy-
namic procedure is used to calculate in space and time, the
constant Cd = Cs2 of the Smagorinsky model (Germano et
al.29). Within this procedure a test filter, denoted by “∧” and 7. Mixed model
typically with filter width 2⌬, is used to calculate the re-
The mixed model (MIX) was designed to overcome the
solved turbulent stresses’ tensor,
difficulties of the scale-similarity model described above. It
Lij = 共u⬍ ⬍ ∧ ⬍ ∧ ⬍ ∧
i u j 兲 − 共ui 兲 共u j 兲 共17兲 consists in adding a dissipative Smagorinsky term to the
original scale-similarity model (Bardina, Ferziger, and
and
Reynolds31),
M ij = k2共⌬x兲2兩S⬍兩∧共S⬍ ∧ ⬍ ⬍ ∧
ij 兲 − 共⌬x兲 共兩S 兩Sij 兲 .
2
共18兲
␶ij = ␣关共u⬍ ⬍ ⬍ ⬍ ⬍ ⬍ ⬍ 2 ⬍ ⬍
i u j 兲 − 共ui 兲 共u j 兲 兴 − 2共Cs⌬兲 兩S 兩Sij . 共22兲
Here k = 冑5 corresponds to a test filter of ⌬ = 2⌬x (Vreman,
Here we took ␣ = 1 as all studies seem to indicate ␣ ⬇ 1 as
Geurts, and Kuerten8). The constant Cd is then given by
the best choice (Meneveau and Katz3). Moreover, with ␣
具M ijLij典 = 1 the model is Galilean invariant (Speziale33).
Cd共xជ ,t兲 = . 共19兲
具M ij M ij典
The brackets indicate averaging along the homogeneous flow
directions (x and z in the present temporal plane jet simula- III. COMPUTATIONAL DETAILS
tions). If, during the calculation process, Cd共xជ , t兲 ⬍ 0 some- A. Numerical method
where, the subgrid-scale model becomes a source of kinetic
energy instead of dissipating it. It was thought in the early The numerical code used in the present simulations is a
days of dynamic modeling to be a very interesting feature of standard pseudospectral code (collocation method34) in
the model since it seemed that the model could provide back- which the temporal advancement is made with an explicit
scatter (Piomelli and Chasnov30). It was later found that it third-order Runge–Kutta time stepping scheme
can, in fact, lead to numerical problems. In practical compu- (Williamson35). All simulations reported in this paper were
tations with the dynamical model Cd is normally set artifi- fully dealised using the 2 / 3 rule (Canuto et al.34).
cially to zero whenever its value given by Eq. (19) is nega-
B. Physical and computational parameters
tive.
Direct and large-eddy simulations of temporal evolving
turbulent plane jets were carried out in this work. Table I lists
5. Scale-similarity model
the details of all the simulations.
In the scale-similarity model (SS) no eddy-viscosity as- The direct numerical simulation (DNS) was carried out
sumption is made. It was developed by Bardina, Ferziger, in order to analyze in detail the physical mechanisms of the
and Reynolds31 to exploit the fact that there is a strong cor- SGS enstrophy dissipation [term VI in Eq. (2)] and to con-
relation between the smallest grid scale and subgrid-scale duct a priori tests on this quantity using several SGS models.
stresses. The model is defined as Although the a priori tests provide many useful and impor-
tant pieces of information, it is generally agreed that they
␶ij = ␣Lij = ␣关共u⬍ ⬍ ⬍ ⬍ ⬍ ⬍ ⬍
i u j 兲 − 共ui 兲 共u j 兲 兴. 共20兲
tend to be too severe. In actual LES the modeled SGS
The constant ␣ was here taken to be equal to 1. Very good stresses tend to adjust to the simulated flow and in general do
correlation was found between the actual and modeled not bring consequences as bad as one would suppose. There-
subgrid-scale stress tensor ␶ij; however, because the model fore, whenever possible, it is always important to analyze
can include backscatter, it was found that this model does not results from actual LES.2 For this purpose, various LES were
dissipate sufficient energy leading to numerical problems. To carried out as a posteriori tests in order to confirm the pre-
overcome this drawback, the scale-similarity model is usu- vious analysis.

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4510 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

TABLE I. Details of the direct and large-eddy simulations.

Simulationsa Subgrid-scale model 共N1 , N2 , N3兲b Resolutionc

DNS (none) 256⫻ 384⫻ 128 ⌬ = 0.02h

SMAG_2dx Smagorinsky 128⫻ 192⫻ 64 ⌬ = 0.04h


SF_2dx Structure function 128⫻ 192⫻ 64 ⌬ = 0.04h
FSF_2dx Filtered structure function 128⫻ 192⫻ 64 ⌬ = 0.04h
DSMAG_2dx Dynamic Smagorinsky 128⫻ 192⫻ 64 ⌬ = 0.04h
SS_2dx Scale similarity 128⫻ 192⫻ 64 ⌬ = 0.04h
GRAD_2dx Gradient 128⫻ 192⫻ 64 ⌬ = 0.04h
MIX_2dx Mixed 128⫻ 192⫻ 64 ⌬ = 0.04h

SMAG_4dx Smagorinsky 64⫻ 96⫻ 32 ⌬ = 0.08h


SF_4dx Structure function 64⫻ 96⫻ 32 ⌬ = 0.08h
FSF_4dx Filtered structure function 64⫻ 96⫻ 32 ⌬ = 0.08h
DSMAG_4dx Dynamic Smagorinsky 64⫻ 96⫻ 32 ⌬ = 0.08h
SS_4dx Scale similarity 64⫻ 96⫻ 32 ⌬ = 0.08h
GRAD_4dx Gradient 64⫻ 96⫻ 32 ⌬ = 0.08h
MIX_4dx Mixed 64⫻ 96⫻ 32 ⌬ = 0.08h

SMAG_8dx Smagorinsky 32⫻ 48⫻ 16 ⌬ = 0.16h


SF_8dx Structure function 32⫻ 48⫻ 16 ⌬ = 0.16h
FSF_8dx Filtered structure function 32⫻ 48⫻ 16 ⌬ = 0.16h
DSMAG_8dx Dynamic Smagorinsky 32⫻ 48⫻ 16 ⌬ = 0.16h
SS_8dx Scale similarity 32⫻ 48⫻ 16 ⌬ = 0.16h
GRAD_8dx Gradient 32⫻ 48⫻ 16 ⌬ = 0.16h
MIX_8dx Mixed 32⫻ 48⫻ 16 ⌬ = 0.16h

SMAG_16dx Smagorinsky 16⫻ 24⫻ 8 ⌬ = 0.32h


SF_16dx Structure function 16⫻ 24⫻ 8 ⌬ = 0.32h
FSF_16dx Filtered structure function 16⫻ 24⫻ 8 ⌬ = 0.32h
DSMAG_16dx Dynamic Smagorinsky 16⫻ 24⫻ 8 ⌬ = 0.32h
SS_16dx Scale similarity 16⫻ 24⫻ 8 ⌬ = 0.32h
GRAD_16dx Gradient 10⫻ 24⫻ 8 ⌬ = 0.32h
MIX_16dx Mixed 16⫻ 24⫻ 8 ⌬ = 0.32h
a
The symbols “mdx” with m = 2, 4, 8, 16 (e.g., SMAG_2dx) mean that the LES corresponds to the case where the implicit (cutoff) LES filter is placed at
kcm = ␲ / m⌬x.
b
N1 , N2, and N3 are the number of collocation points along the x , y, and z directions, respectively.
c
In all simulations we have ⌬ = ⌬x = ⌬y = ⌬z.

The DNS was started using, as initial condition, a h / ␪0 ratio and initial amplitude noise, although favoring a
hyperbolic-tangent velocity profile (da Silva and Métais6), faster transition to turbulence, does not affect the dynamics

U共x,y,z兲 =
U1 + U2 U1 − U2
2

2
tanh
h
4␪0
1−冋 冉
2兩y兩
h
冊册 ,
of the self-similar, fully developed turbulent state (Stanley
and Sarkar,36 Stanley, Sarkar, and Mellado,37 da Silva and
Métais6).
共23兲 The computational domain had the same size in all simu-
lations. It extends to 共Lx , Ly , Lz兲 = 共5.5h , 8h , 2.6h兲 along the
to which a relatively high amplitude white noise was added
streamwise 共x兲, lateral 共y兲, and spanwise 共z兲 directions, re-
(8%) to speed up the transition mechanism and let the flow
spectively. Moreover, all the simulations were carried out
reach quickly a fully developed turbulent state. In Eq. (23) ␪0
with an isotropic grid, with ⌬ = ⌬x = ⌬y = ⌬z in all the simu-
is the initial momentum thickness, h is the jet’s inlet slot
lations. See Table I for details.
width, and we used U1 = 1 and U2 = 0. The initial Reynolds
number was Reh = 1500. The ratio h / ␪0 was set to 20 which
C. The filters
gives about 26 collocation points at each shear layer, at the
start of the simulation. This allows a very neat resolution of In this work two filters with several filter widths will be
the initial shear layer. For comparison da Silva and Métais6 used to separate between the large and small scales of mo-
used h / ␪0 = 30 in direct numerical simulations of plane jets tion, through a filtering operation defined by Eq. (3).
and could still get all the details of the primary Kelvin– The emphasis in this work is placed on physical space
Helmholtz instabilities, although using about 13 points per analysis/representation because the goal is to analyze the ef-
shear layer. It is important to stress that the addition of a high fect of the SGS models on the coherent vortices computed in

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4511

LES, and for this purpose it is important to use a localized


filter in the physical space (Liu, Meneveau, and Katz38).
Moreover, most of LES codes use finite differences and finite
volume codes, where the discretization is implicitly associ-
ated with a box filtering operation (Rogallo and Moin39).
Therefore, in most cases we will be using a box or top-hat
filter in which the filter kernel G⌬ is defined as

冦 冧
1 ⌬
if 兩xជ − ␰ជ 兩 ⬍
G⌬共xជ − ␰ជ 兲 = ⌬ 2
0 otherwise.
However, we will also use the cutoff filter because we
will be comparing our filtered DNS data with results from
several LES, which were performed with the present pseu-
dospectral code. The kernel for the spectral cutoff filter is

冦 冧
FIG. 1. Temporal evolution of the maximum vorticity (components) for the

1 if k 艋 kc = plane jet DNS.
Ĝkc共k兲 = ⌬
0 otherwise.
Figures 2(a)–2(d) show the coherent vortices present at
Finally, when computing local correlations with several the self-similar region of the plane jet 共T / Tref ⬇ 46兲, through
flow variables, it is important to use centered filters so that isosurfaces of low pressure and positive Q = 1 / 2共⍀ij⍀ij
the correlation between the two variables is computed using − SijSij兲 (Weiss,40,41 Hunt, Wray, and Moin42).
the same grid location for both variables. This avoids the The larger flow vortices are identified with low pressure
case where noncentered variables make it impossible to com- isosurfaces [Figs. 2(a) and 2(c)] and show the classical span-
pare results from the filtered DNS with the new LES. Since wise rollers present in plane jets. The smaller vortices are
the pseudospectral schemes used here imply limitations on more clearly seen through positive Q isosurfaces [Figs. 2(b)
the number of grid points used for each direction, when us- and 2(d)]. As can be seen, by the time the flow reaches self-
ing the box filter we will have to rely on both centered and similarity most vortices have lost any spatial preferential ori-
noncentered filters. A discussion on this subject is shown in entation, sign of a fully developed turbulent state. We will be
the Appendix. interested in analyzing the influence of several SGS models
Finally, to analyze the influence of the filter width in the on this type of structures.
results, we had to use several different filter widths for the At this stage the Reynolds number based on the stream-
box and cutoff filters. For the former we used kcm = ␲ / m⌬x wise Taylor microscale varies across the shear layer between
with m = 2, 4, 8, 16, while in the latter we used ⌬n = n⌬x with Re␭x = 90 and 100, which shows that the present DNS is rep-
n = 2, 4, 8, 16 (noncentered filters) and n = 3, 5, 9, 17 (cen- resentative of a fully developed turbulent state.
tered filters). Figures 3(a) and 3(b) show one point statistics for the
plane jet DNS compared with experimental results from Gut-
IV. DNS ASSESSMENT mark and Wygnansky,43 Ramparian and Chandrasekhara,44
Thomas and Prakash,45 and the direct numerical simulations
Figure 1 shows the evolution of the maximum box vor- from Stanley, Sarkar, and Mellado37 and da Silva and
ticity components, Métais.6 The details of the data gathered in Fig. 3 are dis-
played in Table II. For all the cases the profiles at adjacent
⍀iM 共t兲 = max兵␻2i 共x,y,z,t兲其 共24兲
stations collapsed, indicating self-similarity. As can be seen,
for the plane jet DNS, with i = 1, 2, 3, where ␻i is the vor- although the scatter between the experiments and computa-
ticity vector. The simulation is run until T / Tref ⬇ 110, where tions is high, the mean streamwise velocity profile and lateral
Tref = H / 2共U1 − U2兲. The simulation starts with a rapid Reynolds stresses from the DNS results agree well with the
increase in enstrophy as the flow evolves into fully devel- data available. Similar agreement exists for the streamwise
oped turbulence. This corresponds to the growth of the pri- and spanwise Reynolds stresses.
mary Kelvin–Helmholtz instability followed by other sec- The resolution used in the DNS is ⌬x ⬇ 3␩ to ⌬x ⬇ 4␩,
ondary instabilities also present in spatially growing plane where ␩ is the Kolmogorov microscale, which is similar to
jets (da Silva and Métais6). Transition to turbulence is com- the resolution used in previous direct numerical simulations
plete by about T / Tref ⬇ 15, since from this point, there is no of plane jets (Stanley, Sarkar, and Mellado,37 Akhavan et
significant difference between the vorticity components. En- al.,46 da Silva and Métais6).
strophy reaches a maximum at 20⬍ T / Tref ⬍ 30 and then de- Finally, Figs. 4(a) and 4(b) show the spatial, two-
cays smoothly until the end of the simulation. Self-similarity dimensional kinetic energy and enstrophy spectra of the
in the first- and second-order moments is attained at about three-dimensional velocity field from the center of the shear
T / Tref ⬇ 30 and maintained until T / Tref ⬇ 60, where confine- layer 共y / H = 0兲 at the self-similar regime 共T / Tref ⬇ 46兲. The
ment effects begin to show in the one point statistics. figures show also the width of the filters used later in the

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4512 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

FIG. 2. Isosurfaces of low pressure (a,c) and Q ⬎ 0 (b,d) for the plane jet DNS at T / Tref ⬇ 46. (a,b) Side view; (c,d) top view. The thresholds are Q
= 15共U1 / h兲2 and p = −0.05␳U21, respectively.

analysis (see below). The kinetic energy spectrum has a fully developed turbulent plane jet. This completes the vali-
−5 / 3 region over about one decade followed by a smooth dation of the turbulent plane jet DNS.
decay at high wave numbers, which shows that the resolution
used in the present simulation is indeed sufficient to accu- V. RESOLVED ENSTROPHY DYNAMICS
rately represent the small-scale motions. The enstrophy spec-
A. Data bank description
trum attains its peak between k2D ⬇ 20 which corresponds to
the start of the dissipation region. The aim of this section is to analyze the enstrophy SGS
These results show that the present DNS is both accu- dissipation in a temporal simulation of a turbulent plane jet,
rate, at the large and small scales, and representative of a but before starting with the analysis of the DNS results, it is

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4513

FIG. 3. Mean velocity profile and lateral Reynolds stresses from the plane jet DNS in the self-similar region 共T / Tref ⬇ 46兲. The present results are compared
with the experimental measurements from Gutmark and Wygnansky (Ref. 43), Ramparian and Chandrasekhara (Ref. 44), Thomas and Prakash (Ref. 45) and
with the direct numerical simulations from Stanley, Sarkar, and Mellado (Ref. 37) and da Silva and Métais (Ref. 6). (a) Mean streamwise velocity profile; (b)
lateral Reynolds stresses profile. ␦U is the half-width of the jet.

important to describe briefly, how the DNS data bank was


created. It consists of 20 instantaneous fields from the plane
jet DNS taken from T / Tref ⬇ 41 to T / Tref ⬇ 50, i.e., from the
self-similar regime. All the mean profiles were computed us-
共25兲
ing both averaging in the two homogeneous directions (x and
z) and averaging over the 20 instantaneous fields. Other In Eq. (25) XVSGS = −2␶ijS⬍ ij ⬎ 0 represents kinetic
quantities such as root-mean square (rms), correlations, prob- energy transfer from grid into the subgrid scales (forward
ability density functions (PDFs), and joint probability den- scatter) while XVSGS = −2␶ijS⬍ ij ⬍ 0 represents transfer from
sity functions (JPDFs) were computed using data from the subgrid scales into grid scales (backward scatter).
centerline plane and the two upper and two lower grids To assess the degree of equilibrium in the time interval
closer to it (i.e., planes y / H = −0.04; −0.02; 0; 0.02; 0.04). used to build the DNS data bank Fig. 5 displays mean pro-
The procedure is carried out in much the same way as in files of terms XIIISGS and XVSGS for the present simulation.
da Silva and Métais.6 The figure shows that as the filter width decreases 共⌬ → 0兲,
An interesting issue that can be analyzed is the subgrid- more energy flows from the grid into the subgrid scales
scales global equilibrium hypothesis. It states that all the 共XVSGS兲, and the more is dissipated by molecular viscosity
kinetic energy transferred from the grid scales into subgrid XIIISGS, but for ⌬ / ⌬x = 2, 4, 8 the terms balance quite rea-
scales is dissipated by the molecular viscosity (da Silva and sonably. The small lack of balance for the case ⌬ / ⌬x = 16
Métais6). Starting from the transport equation for the near the centerline may imply that this filter is placed at the
subgrid-scales kinetic energy, the global equilibrium can be large scales of the flow, i.e., significant kinetic energy is
expressed by being produced at scales smaller than ⌬ / ⌬x = 16. This shows

TABLE II. Details of the experimental and numerical results compared in Fig. 3.

Source Reh h/␪ Initial condition Data collected at

Gutmark and Wygnanskya 30 000 … Laminar flow x / h = 140


Ramparian and Chandrasekharab 1 600 67 Laminar flow x / h = 40
Thomas and Prakashc 8 000 29 Laminar flow x / h = 14

Stanley, Sarkar, and Melladod 3 000 20 White noise x/h=8


da Silva and Métaise 3 000 20 White noise x / h = 10
a
Reference 43.
b
Reference 44.
c
Reference 45.
d
Reference 37.
e
Reference 6.

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4514 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

FIG. 4. Two-dimensional (spatial) kinetic energy and enstrophy spectra from the plane jet DNS in the center of the shear layer 共y / H = 0兲 at T / Tref ⬇ 46. (a)
Kinetic energy spectrum; (b) enstrophy spectrum. The figures show also the position of several filter widths used in this work. Noncentered filters: ⌬ / ⌬x
= 2, 4, 8, 16. Centered filters: ⌬ / ⌬x = 3, 5, 9, 17.

that there is some degree of equilibrium even in the smallest We start analyzing the filtered enstrophy dynamics by
scales of motion which gives us further confidence in the use plotting mean profiles of the dominating terms from Eq. (2)
of the DNS data bank. in Figs. 6(a)–6(d). It was observed that in the mean, terms II
(filtered enstrophy convection) and IV (viscous enstrophy
B. Mean enstrophy SGS dissipation diffusion) are negligible and were therefore removed from
To analyze the dynamics of the enstrophy SGS dissipa- the figures for clarity. Thus, the evolution of the enstrophy is
tion (VI) we compare this term with the other mechanisms mainly governed by its production (term III), viscous dissi-
also present in the resolved enstrophy transport equation (2), pation (term V), and SGS dissipation (term VI). In the mean,
like the enstrophy production (III) and the enstrophy viscous the SGS enstrophy dissipation (VI) acts as a sink of enstro-
dissipation (V), which were already well documented in phy (always negative), thus the coherent vortices obtained
other flows (e.g., Tsinober, Ortenberg, and Shtilman47 in di- from LES should be subjected to an additional (nonviscous)
rect numerical simulations of homogeneous isotropic turbu- dissipation mechanism.
lence). Figures 6(b)–6(d) show the effect of the filter width on
the enstrophy production (III), enstrophy viscous dissipation
(V), and enstrophy SGS dissipation (VI). Both the enstrophy
production and viscous dissipation decrease with increased
filter width [Figs. 6(b) and 6(c)]. The fact that these terms
decrease so much with only a small increase in filter width
shows that a large amount of enstrophy is being generated at
the smallest scales of motion. The enstrophy SGS dissipation
(VI), on the other hand increases first from ⌬ / ⌬x = 2 to
⌬ / ⌬x = 4, but decreases after from ⌬ / ⌬x = 4 to ⌬ / ⌬x = 16,
although not as fast as the enstrophy viscous dissipation (V).
Comparing the magnitudes of these terms for each filter,
and considering that they should balance for each filter
width, one sees that for the smaller filter width 共⌬ / ⌬x = 2兲,
the enstrophy production 共III2兲 is mainly balanced by the
viscous dissipation 共V2兲, while the SGS dissipation 共VI2兲 is
negligible. For ⌬ / ⌬x = 4 the enstrophy viscous dissipation
共V4兲 and enstrophy SGS dissipation 共VI4兲 have more or less
the same importance, but for ⌬ / ⌬x = 8, however, the enstro-
phy viscous dissipation 共V8兲 becomes less important than the
FIG. 5. Profiles of the dominating terms in the SGS kinetic energy equation
[Eq. (25)]. The terms were computed using no filtering (0) and using the box enstrophy SGS dissipation 共VI8兲 which dissipates most of the
filter with three filter widths (⌬ / ⌬x = 2, 4, 8, 16). enstrophy produced. For ⌬ / ⌬x = 16 the enstrophy viscous

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4515

FIG. 6. Profiles of the dominating terms in the enstrophy transport equation (2) using no filtering (0) and using the box filter with three different filter widths
(⌬ / ⌬x = 2, 4, 8, 16). (a) Enstrophy production (III) and viscous dissipation (V); (b) filtered enstrophy production with ⌬ / ⌬x = 0, 2, 4, 8, 16; (c) filtered
enstrophy viscous dissipation with ⌬ / ⌬x = 0, 2, 4, 8, 16; (d) filtered enstrophy SGS dissipation with ⌬ / ⌬x = 0, 2, 4, 8, 16.

dissipation 共V16兲 is virtually zero and all the enstrophy pro- To complete the characterization of the dominating terms
duction is balanced by the enstrophy SGS dissipation 共VI16兲. from the resolved enstrophy transport equation, and their de-
Figure 7 shows the rms of the dominating terms from the pendence on the filter width, we integrate the mean profiles
resolved enstrophy transport equation with several filter of the enstrophy viscous dissipation (V) and enstrophy SGS
widths, in order to compare the mechanisms of viscous and dissipation (VI) shown in Fig. 6. The evolution of these
SGS dissipation (V and VI). Looking at the rms, which is a quantities, defined by
measure of the local intensity of each term, we see that for
⌬ / ⌬x = 2 the local intensity of both terms (V and VI) is com-
parable, but for ⌬ / ⌬x 艌 4 the SGS dissipation (VI) is much
Vtotal = 冕 +Ly/2

−Ly/2
冓 ␯
⳵ ⍀⬍ ⬍
i ⳵ ⍀i
⳵ xj ⳵ xj

共y兲 dy 共26兲

more active and similar to the enstrophy production (III)


and

冓 冔
while the viscous dissipation (V) is negligible in comparison.
In short, as the filter moves into larger scales, i.e., more into
the inertial and further away from the dissipative range, the VItotal = 冕 +Ly/2

−Ly/2
␧ijk⍀⬍
i
⳵2␶kp
⳵ x j ⳵ xp
共y兲 dy, 共27兲
SGS enstrophy dissipation (VI) becomes more and more im-
portant, while the effects of molecular viscosity become neg- is shown in Fig. 8. As the filter width increases the impor-
ligible. tance of the viscous dissipation (V) decreases monotonically,

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4516 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

FIG. 7. Rms of the enstrophy vorticity production (III), enstrophy viscous FIG. 9. PDFs of relevant terms from the filtered enstrophy transport equa-
dissipation (V), and enstrophy SGS dissipation (VI) at the far field of the tion (2), using the box filter with ⌬ / ⌬x = 4. III, enstrophy production; V,
plane jet DNS. The statistics were obtained through the application of the enstrophy viscous dissipation; VI, enstrophy SGS dissipation. For each PDF
box filter with filter widths ⌬ / ⌬x = 0, 2, 4, 8, 16. ⌬ = 0 means that no filter computation, all variables were removed from their mean values and non-
was applied to the terms. dimensionalized by their respective variances.

whereas the enstrophy SGS dissipation (VI) increases first


and then decreases slightly. ⌬ / ⌬x ⬇ 5 marks the point where will show only the ones obtained with the filters ⌬ / ⌬x = 2
the enstrophy SGS dissipation becomes more important than (dissipative range filter) and ⌬ / ⌬x = 8 (inertial range filter).
the viscous dissipation. The ratio VItotal / Vtotal seems to be
tending to a constant and slow growth rate after this. C. Probability density functions and spatial-temporal
The change in the relative importance between the en- correlations
strophy viscous dissipation and SGS dissipation can be more
clearly understood by examining again Figs. 4(a) and 4(b) To characterize the local (in the physical space) behavior
which show the location of the different filter widths in the of these terms, Fig. 9 shows their PDFs using a box filter
kinetic energy spectrum. Here one can see that the filter with ⌬ / ⌬x = 4. It was observed that the width of the filter has
⌬ / ⌬x = 2 is at the middle of the dissipative region of the a very small influence in the shape of these PDFs. Notice
spectrum, where viscosity effects play a vital role, whereas that in all the single or joint PDFs computed in this paper,
the filter ⌬ / ⌬x = 8 is at the inertial range region, where vis- prior to the computation of the PDF, all variables were re-
cosity effects are negligible. moved from their mean values and were nondimensionalized
Although we used all the four filter widths in this work, by their respective variances.
due to limitations of space, whenever a comparison with the The sign of the absolute values obtained by each term
results obtained with several filter widths is necessary we are listed in Table III, e.g., it was verified that the enstrophy
viscous dissipation (V) exhibited always negative values,
whereas the enstrophy production (III) and enstrophy SGS
dissipation (VI) could be either positive or negative.
Therefore, Fig. 9 shows that although the enstrophy pro-
duction (III) can represent either production (positive) or de-
struction (negative) of resolved enstrophy, production events
dominate, making the overall contribution of this term posi-
tive, as described above and in agreement with, e.g., Tsi-
nober, Ortenberg, and Shtilman,47 Galati and Tsinober.48
Also, the viscous enstrophy dissipation (V) is always nega-
tive and the term always contributes to the dissipation of the
resolved enstrophy. For the enstrophy SGS dissipation (VI)
we see that it can either represent a forward enstrophy cas-
cade (negative) or a backward enstrophy cascade (positive),
but, as we saw before, in the mean, the dissipation mecha-
nism (forward scatter) prevails.
Table III shows the correlation coefficients between sev-
eral terms from Eqs. (2) and (25) for filter widths ⌬ / ⌬x = 3
and ⌬ / ⌬x = 9. The table shows that, with few exceptions, as
FIG. 8. Vtotal and VItotal for filter widths ⌬ / ⌬x = 0, 2, 4, 8, 16 using a box the filter width changed from ⌬ / ⌬x = 3 to 9 the maximum
filter. variation of the correlation coefficients was ⬍0 . 1. For sim-

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4517

TABLE III. Correlation coefficients between several terms from Eqs. (2) and (25) obtained with a box filter
with ⌬ / ⌬x = 3 and ⌬ / ⌬x = 9. 兩S⬍兩 is the filtered rate of strain and 兩⍀⬍兩 the filtered vorticity norm.

Correlation ⌬ / ⌬x 兩⍀⬍兩 兩S⬍兩 III V VI XIIISGS XVSGS

Signa P P NP N NP N NP

III 3 +0.48 +0.37 ⫹1.00 −0.36 −0.62 −0.43 +0.74


9 +0.45 +0.32 ⫹1.00 −0.25 −0.54 −0.41 +0.69

V 3 −0.48 −0.46 −0.36 ⫹1.00 +0.22 +0.94 −0.39


9 −0.42 −0.31 −0.25 ⫹1.00 +0.17 +0.74 −0.31

VI 3 −0.35 −0.33 −0.62 +0.22 ⫹1.00 +0.34 −0.54


9 −0.36 −0.24 −0.54 +0.17 ⫹1.00 +0.34 −0.39
a
Sign: P, always positive; N, always negative; NP, can be negative or positive.

plicity, we will discuss the correlation coefficients for the present in term VI [term VIII in Eq. (4)] and consider only
case ⌬ / ⌬x = 3, but the same conclusions are valid for the enstrophy SGS transfer, term VII in Eq. (4). We obtain
⌬ / ⌬x = 9 (see Table III). Corr共III, VII兲 = −0.47 for ⌬ / ⌬x = 3 and Corr共III, VII兲 = −0.35
The correlations show that the dominating terms have for ⌬ / ⌬x = 9 against Corr共III, VI兲 = −0.62 and −0.54 for
some correlation both with the regions of high filtered vor- ⌬ / ⌬x = 3 and 9, respectively. Thus, the approximation,
ticity and high filtered deformation rate [note that these
quantities are themselves correlated, Corr共兩S兩 , 兩⍀兩兲 = 0.63].
However, the enstrophy production (III), viscous dissipation 共29兲
(V), and SGS dissipation (VI) exist preferentially in the re-
gions of high vorticity, e.g., Corr共兩⍀⬍兩 , III兲 ⬎ Corr共兩S⬍兩 , III兲. is not as good as that of Eq. (28). Therefore, care should be
The joint PDF (JPDF) of these correlations is shown in taken when using Eq. (29) as a starting point for a turbulence
Figs. 10(a)–10(c). Again, before computing the JPDFs all model for the vorticity transport equation as in, e.g., Mans-
terms were removed from their mean values and were non- field, Knio, and Meneveau.18
dimensionalized by their variances. It was observed that Moreover, we note that, as expected, there is a very high
there are only small differences when comparing the joint correlation between the kinetic energy transfer into the
PDF of the terms and the vorticity norm or the deformation subgrid scales 共XVSGS兲 and the enstrophy production
rate. Note that the correlation between the enstrophy produc- (III), Corr共III, XVSGS兲 = + 0.74, which supports the common
tion and the vorticity comes mainly from III⬎ 0 (production) conception that production of vorticity is associated with en-
and the correlation between the enstrophy SGS dissipation ergy cascading into smaller scales. The JPDF of terms III and
and the vorticity comes mainly from VI⬍ 0 (enstrophy for- XVSGS in Fig. 11 shows that this correlation comes both from
ward scatter). III, XVSGS ⬎ 0 and III, XVSGS ⬍ 0.
It is interesting to notice that the two enstrophy dissipa- Another interesting observation is the high correlation
tion mechanisms (viscous V and inviscid VI) act at quite between the viscous dissipation of enstrophy and SGS ki-
different locations. First, the two mechanisms are not corre- netic energy Corr共V , XIIISGS兲 = + 0.94. Unlike the other
lated 关Corr共V , VI兲 = + 0.22兴 and, while the enstrophy viscous terms in Eqs. (2) and (25), these terms are always negative.
dissipation (V) is also quite uncorrelated with the enstrophy They play a similar role in the dynamics of the resolved
production (III) 关Corr共III, V兲 = −0.36兴, the enstrophy SGS enstrophy and the SGS kinetic energy associated with the
dissipation (VI) does have some reasonable correlation “natural” sink caused by molecular viscosity. Figure 12
关Corr共III, VI兲 = −0.62兴. Thus, there is some kind of “local shows the JPDF of these terms and it seems that their corre-
equilibrium” between enstrophy creation (III) and dissipation lation is equally large either from Corr共V , XIIISGS兲
by the subgrid scales (VI) (in reality enstrophy transfer into Ⰶ 0 , Corr共V , XIIISGS兲 ⬍ 0, or Corr共V , XIIISGS兲 ⬇ 0. This
smaller scales of motion). This is particularly true in the demonstrates that the ends of the energy and enstrophy cas-
inertial range where, as we saw before, the viscous enstrophy cades are spatially correlated.
dissipation is negligible and enstrophy production is mostly Finally, we see that the enstrophy SGS dissipation (VI)
balanced by the enstrophy SGS dissipation. Thus, is more correlated with the kinetic energy SGS transfer
共XVSGS兲 than the kinetic energy viscous SGS dissipation
共XIIISGS兲, with Corr共VI, XVSGS兲 = −0.54 and
Corr共VI, XIII 兲 = + 0.34, respectively. Their JPDFs are
SGS

shown in Figs. 13(a) and 13(b). It is interesting to note that


共28兲
the correlation between the enstrophy SGS dissipation (VI)
where the above relation holds locally (i.e., at each instant and kinetic energy transfer XVSGS comes mainly from VI⬍ 0
and spatial location) as well as globally, when averaged over (enstrophy forward scatter) and XVSGS ⬎ 0 (energy forward
the entire flow field. Surprisingly the correlation decreases scatter), which shows that energy and enstrophy cascades are
slightly when we discard the diffusion mechanism also correlated when the transfer is done from the large into the

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4518 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

FIG. 10. Joint PDF of dominating terms from Eq. (2) and the (fluctuating) vorticity norm. (a) Enstrophy production (III); (b) viscous dissipation (V); (c)
enstrophy SGS dissipation (VI). Filtering was done using the box filter with ⌬ / ⌬x = 3. For each JPDF computation, all variables were removed from their
mean values and nondimensionalized by their respective variances.

small scales (forward energy and enstrophy cascades) more


than when it is from small to large scales (backward energy
and enstrophy cascades).

VI. A PRIORI TESTS: MODELED ENSTROPHY SGS


DISSIPATION

This section will assess the behavior of the enstrophy


SGS dissipation when modeled a priori by several SGS
models. The results are compared with the “real” enstrophy
SGS dissipation obtained from the filtered DNS data.

A. Mean profiles of enstrophy SGS dissipation


We start with the mean profiles of real and modeled
enstrophy SGS dissipation (VI) shown in Figs. 14(a) and
FIG. 11. Joint PDF of kinetic energy transfer 共XVSGS兲 and enstrophy pro-
14(b) using the box filter with two filter widths: ⌬ / ⌬x = 2 and
duction (III). Filtering was done using the box filter with ⌬ / ⌬x = 3. For the
8. JPDF computation, both variables were removed from their mean values and
For some models, e.g., structure function and gradient, nondimensionalized by their respective variances.

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4519

ter widths ⌬ / ⌬x = 4 and 16 (not shown), in particular, the


great sensibility of the structure function model to the filter
size.
It is interesting to note that, in both cases (⌬ in the
dissipative or inertial range), the gradient and mixed models
give quite similar results. The same happens with the dy-
namic Smagorinsky and scale-similarity models. Finally,
when the filter is placed in the inertial range (arguably where
it should be in a good LES) the structure function and scale-
similarity models give almost identical results. However it
seems that the structure function model is more sensitive
than the other models to variations in filter width.
Although as stressed above, we are mainly interested in
analyzing the enstrophy SGS dissipation in the physical
space, it is always interesting to see the results obtained us-
FIG. 12. Joint PDF of viscous SGS kinetic energy dissipation 共XIIISGS兲 and ing a spectral cutoff filter. One reason for this is to allow the
enstrophy viscous dissipation (V). Filtering was done using the box filter
with ⌬ / ⌬x = 3. For the JPDF computation, both variables were removed
comparison of the a priori tests with the a posteriori results
from their mean values and nondimensionalized by their respective obtained with LES from the present pseudospectral code,
variances. since it makes no sense to compare a priori results obtained
with a box filter with LES results coming from a pseu-
dospectral code.
the relative importance of the modeled to the real enstrophy For this purpose, Figs. 15(a) and 15(b) show the mean
SGS dissipation changes when the filter is placed in the dis- profiles of real and modeled enstrophy SGS dissipation ob-
sipative 共⌬ / ⌬x = 2兲 or in the inertial range region 共⌬ / ⌬x tained with a spectral cutoff filter with kc2 = ␲ / 2⌬x and kc8
= 8兲. Indeed, for ⌬ / ⌬x = 2 the structure function model dissi- = ␲ / 8⌬x, respectively.
pates too much enstrophy, while the gradient model dissi- In general, the same hierarchy of model dissipation is
pates a slightly more than the real enstrophy SGS dissipa- maintained, i.e., the filtered structure function and dynamic
tion. But for ⌬ / ⌬x = 8, both models dissipate less enstrophy Smagorinsky models dissipate the least amount of enstrophy,
then the real enstrophy SGS dissipation. Apart from these while the Smagorinsky and structure function models are the
models, the main trend is more or less the same as one goes ones that cause the highest dissipation if in the dissipative
from ⌬ / ⌬x = 2 to ⌬ / ⌬x = 8, namely, the Smagorinsky model range, while the performance of the latter greatly improves in
always dissipates too much enstrophy, and the dynamic Sma- the inertial range. We do not display the results for the scale-
gorinsky and filtered structure function models are the ones similarity and mixed models since the Leonard stresses be-
that dissipate less. The same conclusions were seen with fil- come zero when using a cutoff filter.

FIG. 13. Joint PDF of enstrophy SGS dissipation (VI) with the (a) kinetic energy SGS transfer 共XVSGS兲; (b) viscous SGS kinetic energy dissipation 共XIIISGS兲.
Filtering was done using the box filter with ⌬ / ⌬x = 3. For each JPDF computation, all variables were removed from their mean values and nondimensionalized
by their respective variances.

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4520 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

FIG. 14. Mean profiles of real and modeled enstrophy SGS dissipation (VI) using several SGS models. SMAG, Smagorinsky; SF, structure function; FSF,
filtered structure function; DSMAG, dynamic Smagorinsky; GRAD, gradient; SS, scale similarity; MIX, mixed. Filtering was done with a box filter with (a)
⌬ / ⌬x = 2; (b) ⌬ / ⌬x = 8.

B. Spectra of enstrophy SGS dissipation The 2D enstrophy SGS dissipation spectra is defined by
Next we analyze the wave number dependence of the → → → →
enstrophy SGS dissipation. For that purpose we define the ⬘ ,t兲tˆ6共K2D,t兲典 = T6共K2D,t兲␦共K2D + K2D
具tˆ6共K2D ⬘ 兲 共32兲
(direct) 2D Fourier transform of the enstrophy SGS dissipa-
tion at the center of the shear layer, and relates to the variance of the enstrophy SGS dissipation

冕冕
⬁ ⬁ through
1
tˆ6共k1,k3,t兲 = t6共x,y = 0,z,t兲e−i共xk1+zk3兲dk1 dk3 ,
共2␲兲2 −⬁ −⬁

共30兲
1 2
具t⬘ 共y = 0,t兲典 =
2 6
冕 0

T6共k2D,t兲dk2D . 共33兲

where Figures 16 and 17 display the (2D spatial) enstrophy

冏 冏
SGS dissipation spectra T6共k2D , t兲, using box and cutoff fil-
⳵2␶kp
t6共x,y = 0,z,t兲 = − ␧ijk⍀⬍ . 共31兲 ters, respectively. For all cases, the differences between the
i
⳵ x j ⳵ xp y=0 real and modeled enstrophy SGS dissipation spectra extends

FIG. 15. Mean profiles of real and modeled enstrophy SGS dissipation (VI) using several SGS models. SMAG, Smagorinsky; SF, structure function; FSF,
filtered structure function; DSMAG, dynamic Smagorinsky; GRAD, gradient. Filtering was done with a spectral-cutoff filter with (a) kc2 = ␲ / 2⌬x; (b) kc8
= ␲ / 8⌬x. Note that using the cutoff filter the scale-similar enstrophy SGS dissipation is zero and the mixed model turns into the Smagorinsky model.

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4521

FIG. 16. Modeled enstrophy SGS dissipation (2D spatial) spectra from the plane jet DNS in the the center of the shear layer 共y / H = 0兲 using the box filter with
(a) ⌬ / ⌬x = 2; (b) ⌬ / ⌬x = 8. SMAG, Smagorinsky; SF, structure function; FSF, filtered structure function; DSMAG, dynamic Smagorinsky; GRAD, gradient;
SS, scale similarity; MIX, mixed.

through the entire wave number spectra. In terms of hierar- real and modeled subgrid-scale stresses is also shown. As in
chy between the models, the figures show that [in light of Eq. Table III the influence of the filter width in the correlations,
(33)], 具t6共x , y = 0 , z , t兲典 should not be very different from as one goes from ⌬ / ⌬x = 3 to 9 amounts to less than 0.1 in
具t6⬘2共x , y = 0 , z , t兲典. For instance, for the box filter at ⌬ / ⌬x most cases. Therefore, in the discussion we will cite the re-
= 2, the Smagorinsky and structure function models are the sults for ⌬ / ⌬x = 3 only, although the conclusions are the
ones that dissipate more enstrophy [see Fig. 14(a)]. same when considering ⌬ / ⌬x = 9.
The correlation between real and modeled subgrid-scale
C. Correlations between real and modeled enstrophy stresses is illustrated in Table IV for the cross stress compo-
SGS dissipation
nent ␶12. Other components lead to similar results. As shown
Table IV shows the correlation coefficients between the in the table, and as described previously in several studies,
enstrophy SGS dissipation and a number of other flow vari- e.g., de Borue and Orszag,32 Liu, Meneveau, and Katz,38 the
ables. The correlation between one component 共␶12兲 of the correlation coefficients for the eddy-viscosity models is quite

FIG. 17. Modeled enstrophy SGS dissipation (2D spatial) spectra from the plane jet DNS in the the center of the shear layer 共y / H = 0兲 using the spectral-cutoff
filter with (a) kc2 = ␲ / 2⌬x; (b) kc8 = ␲ / 8⌬x. SMAG, Smagorinsky; SF, structure function; FSF, filtered structure function; DSMAG, dynamic Smagorinsky;
GRAD, gradient. Note that using the cutoff filter the scale-similar enstrophy SGS dissipation is zero and the mixed model turns into the Smagorinsky model.

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4522 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

TABLE IV. Correlation coefficients involving the modeled enstrophy SGS dissipation (VI), enstrophy produc-
tion (III), viscous dissipation (V), filtered vorticity norm 共兩⍀⬍兩兲 and filtered deformation rate 共兩S⬍兩兲, using
several SGS models, for filter widths ⌬ / ⌬x = 3 and 9. The correlation coefficients for one component 共␶12兲 of the
real and modeled subgrid-scale stresses is also shown for comparison, along with the values obtained by de
Borue and Orszag (Ref. 32) and Liu, Meneveu, and Katz (Ref. 38). SMAG, Smagorinsky; SF, structure
function; FSF, filtered structure function; DSMAG, dynamic Smagorinsky; GRAD, gradient; SS, scale similar-
ity; MIX, mixed.

Correlation ⌬ / ⌬x SMAG SF FSF DSMAG GRAD SS MIX

共␶real
12 , ␶12 兲
model
3 +0.20 +0.19 +0.18 +0.19 +0.99 +0.97 +0.38
9 +0.22 +0.22 +0.19 +0.21 +0.91 +0.87 +0.36
Borue and Orszaga b
+0.23 … … … +0.97 … …
Liu, Meneveau, and Katzc b
+0.20 … … … +0.80 +0.75 +0.40

共VIreal , VImodel兲 3 +0.61 +0.53 +0.53 +0.61 +0.97 +0.94 +0.64


9 +0.46 +0.41 +0.21 +0.45 +0.64 +0.77 +0.47

共III, VImodel兲 3 −0.63 −0.58 −0.49 −0.63 −0.66 −0.70 −0.64


9 −0.54 −0.48 −0.24 −0.56 −0.54 −0.66 −0.55

共V , VImodel兲 3 +0.27 +0.14 +0.20 +0.27 +0.25 +0.28 +0.27


9 +0.34 +0.18 +0.20 +0.35 +0.17 +0.20 +0.35

共兩⍀⬍兩 , VImodel兲 3 −0.56 −0.53 −0.42 −0.56 −0.38 −0.43 −0.56


9 −0.66 −0.63 −0.34 −0.63 −0.32 −0.45 −0.66

共兩S⬍兩 , VImodel兲 3 −0.38 −0.30 −0.24 −0.38 −0.35 −0.38 −0.39


9 −0.34 −0.26 −0.10 −0.34 −0.22 −0.27 −0.34
a
Reference 32.
b
As in the present results, the coefficients obtained by Borue and Orszag (Ref. 32) and Liu, Meneveau, and Katz
(Ref. 38) change by a small amount with the width of the filter considered. The values indicated in the table
correspond to the larger filter considered in these works.
c
Reference 38.

low (less than 0.2) whereas the ones obtained with the gra- the models has a very low correlation with the viscous en-
dient and scale-similarity models are very high (more than strophy dissipation, Corr共V , VImodel兲 ⬇ + 0.20, as in the real
0.8). Moreover, the addition of a scale-similarity term to the case where Corr共V , VI兲 = + 0.22. Moreover, for all the mod-
Smagorinsky model increases substantially the correlation els the enstrophy SGS dissipation occurs more in the vortic-
with the real stresses. We have 0.38 against 0.40 in de Borue ity dominated regions than in the deformation rate dominated
and Orszag.32 regions, Corr共兩⍀⬍兩 , VImodel兲 ⬎ Corr共兩S⬍兩 , VImodel兲, also as in
Considering these correlations for the subgrid-scale the real enstrophy SGS dissipation.
stresses it may be surprising to see that all the models lead to
a very good correlation between real and modeled enstrophy D. Probability density functions of enstrophy SGS
SGS dissipation (see Table IV). This can be explained by the dissipation
presence of the filtered vorticity ⍀⬍ i , which of course is
the same for all models, in the definition of term VI. To analyze the issue of statistical behavior we turn to the
As expected, the gradient and scale-similarity models PDFs of the real and modeled enstrophy SGS dissipation
have the highest correlations 关Corr共VIGRAD , VI兲 (VI), which are shown in Fig. 18 for filter width ⌬ / ⌬x = 8.
SS
= + 0.97, Corr共VI , VI兲 = + 0.94兴, due to the high level of We recall that negative/positive values of enstrophy SGS
correlation between the real and modeled subgrid-scale dissipation (VI) represent forward/backward enstrophy cas-
stresses in these models. cade, respectively. Notice that unlike what happens when
It is interesting to see the correlation between the enstro- considering the grid-scale kinetic energy, all models even of
phy production (III) and the modeled enstrophy SGS dissi- the eddy-viscosity type, provide both a forward and back-
pation, since as we saw before, there is some kind of local ward enstrophy cascade.
equilibrium between these mechanisms in a real flow, pro- Concerning the forward enstrophy tails of the PDFs
vided that ⌬ is placed at the inertial range. The correlation 共VI⬍ 0兲, all the models show very similar results and are
between Corr共III, VImodel兲 is very close to the real value of quite close to the real shape of the PDF of real enstrophy
Corr共III, VI兲 = −0.62 for all the models, except for the filtered SGS dissipation (VI), as shown in Fig. 18 for ⌬ / ⌬x = 8. The
structure function model, and this might explain why it dis- same occurs with other filter widths (not shown). It turns out
sipates a very small amount of enstrophy, when compared to that most differences come from the backward enstrophy tail
the other models. 共VI⬎ 0兲 of the PDFs. This part is shown in more detail in
Finally, notice that the enstrophy SGS dissipation for all Figs. 19(a) and 19(b).

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4523

than the eddy-viscosity models. The exception to this is the


filtered structure function model which is also one of the
closest to the real backward enstrophy cascade, as long as the
filtering is not done at very small scales 共⌬ / ⌬x = 2兲.
Finally, Figs. 20(a) and 20(b) show the JPDFs of the real
and modeled enstrophy SGS dissipation for the Smagorinsky
and the gradient models, with ⌬ / ⌬x = 3. The shape of the
JPDFs for the structure function, filtered structure function,
dynamic Smagorinsky, and mixed models (not shown) is
quite similar to the Smagorinsky model. The figure shows
that for these models the (good) correlation between the real
and modeled enstrophy SGS dissipation comes mainly from
the forward scatter part of the PDF (VI⬍ 0 and VImodel ⬍ 0),
and not from its backscatter tail (VI⬎ 0 and VImodel ⬎ 0),
which, as we saw before, does not match the real enstrophy
FIG. 18. Probability density functions (PDFs) of the real and modeled en-
strophy SGS dissipation (VI) using several SGS models, for a box filter with SGS dissipation. In particular, the JPDF for the mixed model
⌬ / ⌬x = 8. SMAG, Smagorinsky; SF, structure function; FSF, filtered struc- (not shown) is surprisingly similar to the one obtained with
ture function; DSMAG, dynamic Smagorinsky; GRAD, gradient; SS, scale the Smagorinsky model, and from that point of view the
similarity; MIX, mixed.
mixed model does not bring a great additional advantage, in
agreement with the observations made earlier with the single
PDFs. For the scale-similarity and gradient models, however,
First, for ⌬ / ⌬x = 2 one sees that the non-eddy-viscosity the correlation between real and modeled enstrophy SGS dis-
models (GRAD,SS) are the ones with a PDF closest to the sipation is very good for both the forward and backscatter
real enstrophy SGS dissipation. All the eddy-viscosity mod-
parts, as shown in Fig. 20(b) for the gradient model. The
els and the mixed model, although providing some backward
scale-similarity has a very similar JPDF (not shown). The
enstrophy transfer, tend to display much less than the real
influence of the filter size on the shape of the JPDFs can be
value. However, for a filter placed at the inertial range
共⌬ / ⌬x = 8兲 the filtered structure function model joins the appreciated in Fig. 21 showing the JPDF for gradient model
backward enstrophy tail of the real enstrophy SGS dissipa- with ⌬ / ⌬x = 9. In agreement with the correlation coefficients
tion and thus becomes also one of the best models in coping documented in Table IV, the correlations tend to decrease
with this real enstrophy backscatter. This shift from the fil- slightly with an increase in the filter width, and that fact is
tered structure function model occurs for ⌬ / ⌬x 艌 4 (not reflected in the JPDF. However, the correlations are still
shown). quite high, and the result that the non-eddy-viscosity models
In conclusion, in general the non-eddy-viscosity models correlate well both in the forward and backward parts of the
tend to provide a more correct backward enstrophy transfer PDFs remains true.

FIG. 19. Probability density functions (PDFs) of the backscatter part of the real and modeled enstrophy SGS dissipation (VI) using several SGS models, for
a box filter with (a) ⌬ / ⌬x = 2; (b) ⌬ / ⌬x = 8. SMAG, Smagorinsky; SF, structure function; FSF, filtered structure function; DSMAG, dynamic Smagorinsky;
GRAD, gradient; SS, scale similarity; MIX, mixed.

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4524 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

FIG. 20. Joint probability density functions (JPDFs) of real and modeled enstrophy SGS dissipation (VI) using the Smagorinsky and the gradient models with
⌬ / ⌬x = 3. (a) Smagorinsky model; (b) gradient model.

E. Relation to kinetic energy SGS dissipation shown) except for some models (SF, GRAD, MIX) for which
To study the relation between the mean enstrophy and some small differences exist. However, even in these cases
energy SGS dissipation Fig. 22 shows mean profiles of real the differences are very small and the overall picture persists.
and modeled kinetic energy SGS dissipation (term XVSGS), This is an interesting result because it indicates that the clas-
with ⌬ / ⌬x = 8. This quantity has been used extensively in a sical energy SGS dissipation a priori tests might be used to
priori tests to assess the performance of subgrid-scale mod- evaluate the capability of other SGS models not included in
els, e.g., Liu, Katz, and Meneveau,49 Shao, Sarkar, and the present work to reproduce the correct vorticity field. It
Pantano.20 For instance, they found that the energy SGS dis- would be interesting to analyze if this similarity of the SGS
sipation obtained with the Smagorinsky model is greater than models ranking from the energy and enstrophy SGS dissipa-
the real, whereas the scale-similarity model dissipates less tion occurs also in other flow configurations, e.g., wall flows.
energy than that obtained with the filtered DNS. The mixed
model, which combines the two models, improves the
results.49 The same occurs with the present case [see Fig.
F. Influence of the mean field gradient
22(b)].
When we compare the SGS models in terms of energy Since the coherent structures from the far field of turbu-
and enstrophy SGS dissipation [Figs. 14(a), 14(b), 22(a), and lent plane jets are similar to the ones found at other free
22(b)], we see that the ranking of the SGS models matches shear flows (e.g., mixing layers and wakes), it is likely that
almost perfectly. The same is true with ⌬ / ⌬x = 4, 16 (not the present results are also valid for any such free shear
flows, provided the flow is at the far field region. However,
to analyze the dependence of the previous results on the par-
ticular flow considered here, it is instructive to isolate the
influence of the mean flow gradient from the enstrophy SGS
dissipation.
To do this, we introduce the Reynolds decomposition of
the resolved vorticity and subgrid-scale stresses,

⬍⬘
⍀⬍ ⬍
i = 具⍀i 典 + ⍀i , 共34兲

␶ij = 具␶ij典 + ␶⬘ij , 共35兲

FIG. 21. Joint probability density functions (JPDFs) of real and modeled
enstrophy SGS dissipation (VI) for the gradient model with ⌬ / ⌬x = 9. into the definition of the enstrophy SGS dissipation,

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4525

FIG. 22. Mean profiles of real and modeled kinetic energy SGS dissipation 共XVSGS兲 using several SGS models. SMAG, Smagorinsky; SF, structure function;
FSF, filtered structure function; DSMAG, dynamic Smagorinsky; GRAD, gradient; SS, scale similarity; MIX, mixed. Filtering was done with a box filter with
(a) ⌬ / ⌬x = 2; (b) ⌬ / ⌬x = 8.

in Table I and the simulations use the same numerical code


as the DNS.

A. Defining the LES grid sizes


The LES were designed to allow the comparison of their
results with the ones from the filtered DNS with filter widths
kcm = ␲ / m⌬x, for m = 2, 4, 8, 16. Therefore, each LES was
共36兲 carried out in a grid with NLES
i = NDNS
i / m, where NLESi is the
number of collocation points along the “i” direction (i = 1, 2,
In Eq. (36) the contribution of the mean gradient, of the 3) for each LES_mdx and NDNS i is the number of points of
resolved vorticity and subgrid-scale stresses, appears through the reference DNS along that direction (see Table I). Finally,
term 具VI⬘典. Term 具VI⬙典 is caused only by the fluctuating before comparing the LES with the filtered DNS, the DNS
fields. Profiles of these two contributions are displayed in data were filtered into the same grid as the LES. As in the
Fig. 23 and prove that term 具VI⬘典 is overshadowed by 具VI⬙典, reference DNS all LES use grids with ⌬ = ⌬x = ⌬y = ⌬z and
which shows that the mean flow gradients play a minor role with the same physical domain, 共Lx , Ly , Lz兲 = 共5.5h , 8h , 2.6h兲.
in the enstrophy cascade process for the present case. This is
certainly not the case during the transition to turbulence in
free shear flows, where the “fluctuating” statistics are likely
to be dominated by the mean flow gradient. Another case
where one would expect the mean flow gradient to be much
more important is in wall flows, where the kinetic energy
SGS dissipation due to mean field gradients is much more
important than the one due to fluctuating fields (Hartel and
Kleiser50). For these cases it would be interesting to analyze
the ranking of the SGS models in the simulation of the
“mean” and fluctuating parts of the enstrophy SGS dissipa-
tion, as in Hartel and Kleiser.50

VII. A POSTERIORI TESTS: LES OF TEMPORAL


PLANE JETS

In this section several LES are carried out using the SGS
models described earlier, in order to analyze a posteriori the
impact of the SGS models on the evolution of the vorticity FIG. 23. Profiles of (real) enstrophy SGS dissipation (VI) due to mean
field. The computational details of the simulations are listed 共具VI⬘典兲 and fluctuating 共具VI⬙典兲 vorticity and subgrid-scale stresses.

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4526 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

lead to abrupt changes in the early flow development due to


nonphysical perturbations in the initial velocity field. The
same occurs for all the other LES.
Since the maximum vorticity is connected with the high
wave number region of the vorticity spectrum, these figures
reveal clearly how much each model dissipates this quantity.
In all LES_2dx, from start to finish, the structure function
and mixed models cause more dissipation, while the dynamic
Smagorinsky and filtered structure function models cause
less. This agrees with the results from the previous analysis.
Notice that no model is able to reproduce correctly the real
maximum streamwise vorticity, however, this must be seen
as an almost impossible thing to achieve since it is connected
with the end of the high wave number range, from where
most errors in the vorticity field arise, as we will see below.
FIG. 24. Maximum box vorticity for the filtered plane jet DNS and the
LES_2dx. The filter applied to the DNS is a cutoff filter with filter width D. Mean and fluctuating vorticity profiles
kc2 = ␲ / 2⌬x. The crosses represent only some instants of the filtered DNS
between 41⬍ T / Tref ⬍ 50 for the DNS while the lines represent all iterations The exact amount of vorticity dissipated by the several
of the LES_2dx between 0 ⬍ T / Tref ⬍ 10.
SGS models at the same instant (T / Tref ⬇ 46 for the DNS and
T / Tref ⬇ 5 for the LES) can be estimated in Figs. 25(a)–25(d)
which show the profiles of mean vorticity norm for the fil-
tered DNS and LES. In all cases the structure function and
B. Initial conditions for LES mixed models dissipate too much vorticity, compared to the
Great care was taken to design these simulations so that other models. For LES_2dx and LES_4dx all the other mod-
an actual comparison between the LES and filtered DNS els are quite close to the real value of the mean vorticity
could be made. The initial condition used for all LES was the norm and, in agreement with what had been predicted by the
same. It consists on a cutoff 共kcm = ␲ / m⌬x兲 filtered DNS ve- a priori tests, the dynamic Smagorinsky and filtered structure
locity field at T / Tref ⬇ 41, i.e., from the fully developed tur- function models are among the ones which lead to the small-
bulent region. Once filtered, the velocity fields were forced est vorticity dissipation. For the LES_8dx and LES_16dx
to satisfy the continuity constraint and were dealised with the none of the SGS models are able to get the “correct” mean
2 / 3 rule. The LES were performed during an interval of vorticity, but as we will see below, the larger flow vortices
⌬Tsim / Tref ⬇ 10. and the small wave number region of the enstrophy spectra
The goal of this approach in generating the initial con- are being well captured. Thus, the error must come from the
dition was to remove the influence of the transition to turbu- amount of vorticity concentrated in the smallest resolved
lence region, where differences in the behavior of the SGS scales of motion, which is high as we saw before, and is
models are particularly acute. Vreman, Geurts, and Kuerten8 caused by the effect of the SGS models on these scales.
compared the structures from several mixing layer LES both Notice that this error is enough to cause a small phase error
after the transition was completed and within the fully devel- in the position of the coherent vortices in LES_16dx as can
oped turbulence region. As remarked by them, some of the be suspected by the location of the vorticity peaks for the
differences in the structures in the LES were caused by the filtered DNS and the LES_16dx [see Fig. 25(d)]. Apart from
different behavior of the SGS models in the transition region
this fact, the ranking of the SGS models is roughly the same
and not in the fully developed region. The procedure adopted
as in LES_2dx and LES_4dx, although the difference among
in the present study has been used in order to focus on the
them is smaller.
behavior of the models in the fully developed turbulent re-
Figure 26 shows the mean vorticity norm variance for
gion alone.
the LES_2dx. As could be expected from the discussion
C. Maximum vorticity evolution above, no model is able to capture the real vorticity fluctua-
tion; however, the ranking of the SGS models is similar to
We start the comparison between the filtered DNS and the one found in the mean vorticity norm profiles. The same
the LES with the temporal evolution of the maximum observation could be made for the other LES.
streamwise vorticity component ⍀xM shown in Fig. 24 for
simulations LES_2dx. The evolution of ⍀xM in all the simu- E. Enstrophy spectra
lations starts smoothly 共T / Tref ⬍ 1兲 and follows the same
trends as the filtered DNS. It is reassuring to observe this The previous results make it possible to quantify the
type of feature in a variable such as this, which is connected total amount of vorticity but do not specify which scales of
with the high wave numbers and therefore sensitive to small motion are involved. It is possible that the error between the
changes in the flow field. This confirms that the restart from real and LES vorticity comes mainly from the small resolved
the filtered DNS data was indeed well prepared and does not scales, and therefore few signs of this appear at the large

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4527

FIG. 25. Profiles of mean vorticity norm from the filtered DNS (at T / Tref ⬇ 46) and the LES (at T / Tref ⬇ 5). (a) LES_2dx; (b) LES_4dx; (c) LES_8dx; (d)
LES_16dx.

scales of the flow. To analyze this issue, Fig. 27 shows the same found in Figs. 25(a)–25(d). In order of increasing en-
enstrophy (2D spatial) spectrum at the centerline 共y / h = 0兲 strophy dissipation we have for LES_2dx, dynamic Smago-
and at the same instant for the filtered DNS and the LES_2dx rinsky, filtered structure function, gradient, scale-similarity,
in classical log-log coordinates. There is almost no difference Smagorinsky, mixed, and structure function. This ranking is
at all in the enstrophy spectra from the filtered DNS and the similar to the one for LES_16dx, although there the differ-
LES_2dx at the smallest wave numbers, which means that ences between the models are smaller.
the large-scale vorticity or the largest flow vortices from any Notice that Figs. 28(a)–28(d) prove that the error be-
LES_2dx matches the real ones perfectly. The error arises at tween the filtered DNS and the LES vorticity variance found
the intermediate and high wave numbers and tends to in- in Fig. 26 comes from the high wave numbers, since this
crease with the wave number. To highlight the differences quantity is related to the enstrophy spectrum through the
between the models Figs. 28(a)–28(d) show the same enstro- relation,
phy spectra in linear-log coordinates for all the LES. As the
implicit LES filter increases the enstrophy spectrum error


begins to affect smaller wave numbers until for LES_16dx ⬁
1
only one wave number matches the filtered DNS results. 具⍀⬘⍀⬘共y = 0,t兲典 = 2
k2D E共k2D,t兲dk2D . 共37兲
Apart from these facts, the hierarchy of the models is the 2 i i 0

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4528 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

in the filtered structure function model is about 10% in


LES_2dx but reaches 70% in LES_4dx. In general the dy-
namic Smagorinsky, gradient, scale similarity, and filtered
structure function models have a smaller enstrophy spectrum
error than the Smagorinsky, structure function, and mixed
models.
Since the interest of any LES ultimately rests on the low
wave number part of the spectrum, it is important to quantify
the “mean” enstrophy spectrum error at the low wave num-
ber range. For this purpose we define the quantity,

ErrD共k2D ⬍ kint兲 = 冏 Dmodel共k2D兲 − Dreal共k2D兲


Dreal共k2D兲

for k2D ⬍ kint , 共40兲
and its rate in time,

FIG. 26. Profiles of vorticity norm variance from the filtered DNS (at
T / Tref ⬇ 46) and the LES_2dx (at T / Tref ⬇ 5).
ErrT,D共k2D ⬍ kint兲 = 冏 Dmodel共k2D兲 − Dreal共k2D兲
Dreal共k2D兲⌬Tsim

for k2D ⬍ kint , 共41兲
F. Vorticity error estimates
where ⌬Tsim = ⌬T / Tref is the (nondimensional) total simu-
To study the wave number dependence of the enstrophy lated time and kint the maximum wave number of interest. In
spectrum error Figs. 29(a)–29(d) plot the quantity, order to be able to compare results from all the LES we

ErrD共k2D兲 = 冏 Dmodel共k2D兲 − Dreal共k2D兲


Dreal共k2D兲
, 冏 共38兲
choose kint = 5, which is smaller than the implicit cutoff filter
used for all the LES.
Table V shows these quantities for all the LES. From this
where D共k2D兲 is the enstrophy (2D spatial) spectrum, point of view also, the structure function and mixed models
D共k2D兲 = k2D
2
E共k2D兲. 共39兲 give the worst results, for instance with 1.2% –1.4% of en-
strophy spectrum error per simulated time Tref in LES_2dx
For all the models the enstrophy spectrum error in- and 11.6%–12% in the LES_16dx. In general, and in agree-
creases from 0% at k2D = 0 to 100% at k2D = Kmax, but some ment with Fig. 29 the best results are obtained with the dy-
models affect lower wave numbers more than others. For namic Smagorinsky, gradient, and filtered structure function
instance, for the LES_2dx the dynamic Smagorinsky model models with only ⬇0.5% –0.8% error in LES_2dx and
has a low enstrophy spectrum error (less than 10%) up until ⬇11.1% –11.2% in the LES_16dx. Notice that the difference
k2D ⬇ 30, something the structure function and mixed models between the models errors decreases substantially as the filter
can only maintain up to k2D ⬇ 10. Also, in the LES_4dx the width increases.
enstrophy error affects the large and intermediate wave num- It is important to keep in mind that, as any LES is run-
bers much sooner than in LES_2dx, e.g., at k2D ⬇ 20 the error ning, the error at the small scales of motion is expected to
“contaminate” larger scales progressively (Métais and
Lesieur51), i.e., it is normal that the error ErrD共k2D兲 shown in
Fig. 29 will increase with the simulated time, for each SGS
model. However, the previous analysis is still useful to show
which models are affected first, as the cutoff wave number
from LES increases.
The evolution of the total error ErrD共k2D ⬍ kint兲 with the
filter width for kint = Kmax is shown in Fig. 30 for all the LES.
The error decreases with the filter width for most models, as
expected, but interestingly, seems to reach a plateau between
⌬ / ⌬x = 2 and 4 for the mixed, structure function, and Sma-
gorinsky models.

G. Coherent vortices
A comparison of the coherent vortices obtained in the
filtered DNS and the LES (in the same grids) is shown in
Figs. 31–33. The large-scale vortices can be seen through
FIG. 27. Enstrophy spectra at the centerline 共y / h = 0兲 for the DNS at isosurfaces of low pressure, while the smaller vortices are
T / Tref ⬇ 46 and the LES_2dx at T / Tref ⬇ 5 in classical log-log coordinates. better visualized with the aid of positive Q (Dubief and

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4529

FIG. 28. Enstrophy spectra at the centerline 共y / h = 0兲 for the DNS at T / Tref ⬇ 46 and all the LES at T / Tref ⬇ 5 in linear-log coordinates. (a) Enstrophy spectra
for DNS / LES_2dx; (b) enstrophy spectra for DNS / LES_4dx; (c) enstrophy spectra for DNS / LES_8dx; (d) enstrophy spectra for DNS / LES_16dx.

Delcayre52). For greater clarity and due to lack of space, only Similar observation were made for the LES_4dx although in
the structures for two models are shown in each figure and this case the “vorticity loss” at the small structures is even
correspond in each case to the ones that dissipate the greatest bigger than in LES_2dx.
and smallest amounts of vorticity. For LES_8dx Figs. 32(a)–32(f) show the coherent struc-
Positive Q isosurfaces for the filtered DNS and LES_2dx tures obtained with the filtered DNS data and the Smagorin-
for the structure function and dynamic Smagorinsky are dis- sky and dynamic Smagorinsky models. The positive Q ⬎ 0
played in Figs. 31(a)–31(c). There are differences between isosurfaces [(a), (c), and (e)] show that with this grid size the
the structures for each case if one looks at the small-scale small-scale vortices are much more dissipated than before.
vortices. It is clear that some small-scale vorticity was lost in Again the dynamic Smagorinsky model gives better results
the structure function model case, but less with the dynamic than the Smagorinsky, but the difference between the two
Smagorinsky, which remains much closer to the one obtained models is small, specially if we compare both cases with the
in filtered DNS field. On the other hand, there is almost no real filtered DNS results. However, the large flow structures
difference at all if one looks at the large flow vortices ob- are still relatively well captured with both models as shown
tained with the low pressure isosurfaces (not shown), and by the isosurfaces of low pressure [(b), (d), and (f)], even if
this fact agrees with what we saw above: the low wave num- some dissipation of these structures can already be noticed
ber region of the spectrum is not affected by the SGS model when compared with the filtered DNS.
employed, as long as the cutoff filter is not placed close to it. Finally, for LES_16dx even the larger flow vortices be-

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4530 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

FIG. 29. Enstrophy spectrum error ErrD共k2D兲 for all the LES as function of the wave number. (a) LES_2dx; (b) LES_4dx; (c) LES_8dx; (d) LES_16dx.

gin to show a significant dissipation when compared with the In summary, we see in these instantaneous snapshots
structures from the filtered DNS (on the same grid), as what the previous statistical analysis has already shown: As
shown in Figs. 33(a)–33(c). Moreover, there is a clear the filter width moves into larger scales (into the inertial
(phase) error in the LES_16dx compared to the DNS. As range) the vorticity error increases, and smaller wave num-
suggested previously by Fig. 25(d) the large rollers from the bers begin to be affected. However, it is likely that, as the
LES are not at the same position as in the DNS. Thus, the simulations proceed, these small-scale errors will tend to
error present at the small scales has contaminated the larger propagate into increasingly large scales.
flow vortices. Some models “contaminate” the lower wave numbers

TABLE V. Error ErrT,D共k2D ⬍ kint兲 between the filtered DNS and the LES at the centerline 共y / h = 0兲 for kint = 5. The error is defined by Eq. (41).

m SMAG_mdx (%) SF_mdx (%) FSF_mdx (%) DSMAG_mdx (%) GRAD_mdx (%) SS_mdx (%) MIX_mdx (%)

2 0.9 1.2 0.7 0.6 0.8 1.1 1.4


4 0.5 11.9 0.6 1.3 0.9 1.0 1.2
8 4.8 5.5 4.5 4.4 4.6 4.3 5.0
16 11.3 12.0 11.2 11.1 11.3 11.5 11.6

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4531

FIG. 30. Enstrophy error ErrT,D共k2D ⬍ kint兲 for kint = Kmax as function of the
LES implicit filter.

faster than others, e.g., for a filter at the dissipative


range region the structure function, mixed, and Smagorinsky
models cause a higher dissipation of vorticity than the
other models. Among those, the dynamic Smagorinsky and
filtered structure function models seem to lead to the best
results, i.e., more representative of the real vorticity field.
For inertial range LES the Smagorinsky and mixed models
are the most dissipative models while the filtered structure
function and the dynamic Smagorinsky remain the least dis-
sipative.

VIII. CONCLUSIONS

In this work, the effect of the SGS models on the vorti-


ces obtained from LES is analyzed using both DNS and LES
of temporal evolving plane jets at the far field, fully devel-
oped turbulence regime. FIG. 31. Isosurfaces of Q ⬎ 0 for the filtered DNS, structure function and
The evolution of the resolved vorticity (or enstrophy) dynamic Smagorinsky models for the LES_2dx at T / Tref = 46 for the DNS
field in a LES is governed by the vorticity production, vis- and T / Tref = 5 for the LES. The threshold is Q = + 15共U1 / h兲2 in the three
figures and the filtered DNS and the LES_2dx are shown in the same mesh.
cous dissipation, and by an extra term that originates from (a) Filtered DNS; (b) SF_2dx; (c) DSMAG_2dx.
the SGS model: the enstrophy SGS dissipation. The influ-
ence of the SGS models on the vortices computed in LES is
described by this term. Analysis of a DNS data bank showed
that when the implicit LES filter is placed at the inertial
range region, the enstrophy viscous dissipation is negligible SGS dissipation but also all of them are able to provide both
and the resolved vorticity is governed mainly by its produc- a forward and backward enstrophy cascade.
tion and SGS dissipation. Although local and instantaneously Several LES were carried out and compared with results
this term can be either negative or positive (and thus repre- from the filtered DNS results and confirmed the results from
sent a forward/backward enstrophy transfer), in the average, the a priori tests. The SGS models affect the high wave
its value is negative and thus the term represents a sink of numbers much more than the low wave numbers. An estima-
resolved vorticity.
tion of the vorticity error per nondimensional time in the
Classical, a priori tests were conducted in the DNS data
LES is derived for each model. In the end, the dynamic Sma-
bank and showed that, in terms of the spatial correlation with
other mechanisms (e.g., vorticity production and viscous dis- gorinsky and filtered structure function models are the ones
sipation), all the models display the correct physical mecha- which minimize this vorticity error.
nisms. On the other hand, in terms of statistical behavior, the Although this work focuses mainly on a physical space
non-eddy-viscosity type models give the best results. How- analysis, its a posteriori tests were done with a pseudospec-
ever, all the models, even those of the eddy-viscosity type, tral code. Future work should try to verify the results from
not only exhibit a good correlation with the real enstrophy the present a priori tests, in LES using physical space codes,

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4532 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

FIG. 32. Isosurfaces of positive Q (a, c, e) and low pressure (b, d, f) for the filtered DNS, Smagorinsky and dynamic Smagorinsky models for the LES_8dx
at T / Tref = 46 for the DNS and T / Tref = 5 for the LES. The threshold is Q = + 10共U1 / h兲2 and p = −0.05␳U21 in the three figures and the filtered DNS and the
LES_8dx are shown in the same mesh. (a,b) Filtered DNS; (c,d) SMAG_8dx; (e,f) DSMAG_8dx.

i.e., using finite difference and finite volume schemes for APPENDIX: CENTERED AND NONCENTERED
spatial discretization. FILTERS
Finally, we think that it would be interesting for future
In physical space, the filtering procedure applied to any
studies to investigate the influence of the SGS models on the
given flow variable ␾共x兲 requires the solution of the integral
life time and vortex core radius of the coherent vortices ob-
(unidimensional case),
tained from LES.

ACKNOWLEDGMENTS ␾
¯ 共x兲 = 冕⍀
␾共x⬘兲G⌬共x − x⬘兲dx⬘ . 共A1兲

The authors would like to thank Dr. Richard Howard for


Using a box filter the filtered value of ␾共x兲 at node i is
his comments on the initial manuscript and an anonymous
referee who suggested two ideas that we explored in Secs.
VI E and VI F. Moreover, Ricardo Reis is acknowledged for
¯⌬ = 1
␾ i

冕 +⌬/2

−⌬/2
␾共x⬘兲dx⬘ . 共A2兲
providing assistance with Debian Linux. This work was sup-
ported by the Ministério da Ciência e Tecnologia under There are a variety of ways to solve this integral. For ex-
SFRH/BPD/11534/2002. ample, if we want to filter with a filter width equal to ⌬

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
Phys. Fluids, Vol. 16, No. 12, December 2004 Effect of subgrid-scale models 4533

same is true when using high order interpolating methods to


compute this integral.
A more correct way of using a true filter width ⌬ = 2⌬x
would be using the actual values of ␾ at two nodes,

¯ 2⌬x = 1
␾ i
2⌬x
冕 +⌬x

−⌬x
␾共x⬘兲dx⬘

1 ␾i+1 + ␾i
= 关␾i+1⌬x + ␾i⌬x兴 = . 共A4兲
2⌬x 2
This method is equivalent to the computation of a mean
value between two nodes involving one control volume each.
This makes for a total control volume width of 2⌬x. This is
the correct way to apply a box filter with a filter width equal
to 2⌬x, but unfortunately the resulting value should be
placed at a nonexisting i − 1 / 2 node location. This does not
suit the purpose of the present work, because in practice we
are interested in analyzing the exact location of some vari-
ables.
To overcome this problem three control volumes were
used according to the formula,

¯ 3⌬x = 1
␾ i
3⌬x
冕 +⌬x

−⌬x
␾共x⬘兲dx⬘

1
= 关␾i+1⌬x + ␾i⌬x + ␾i−1⌬x兴
3⌬x
␾i+1 + ␾i + ␾i−1
= . 共A5兲
3
This formula represents an exact box filter centered on node
i. The filter width corresponding to this formula is ⌬ = 3⌬x,
because it involves three control volumes with width ⌬x,
each.
1
M. Lesieur and O. Métais, “New trends in large-eddy simulations of tur-
FIG. 33. Isosurfaces of low pressure for the filtered DNS, Smagorinsky and bulence,” Annu. Rev. Fluid Mech. 28, 45 (1999).
2
dynamic Smagorinsky models for the LES_16dx at T / Tref = 46 for the DNS C. Meneveau and J. Katz, “Scale invariance and turbulence models for
and T / Tref = 5 for the LES. The threshold is p = −0.05␳U21 in the three figures large-eddy simulation,” Annu. Rev. Fluid Mech. 32, 1 (2000).
3
and the filtered DNS and the LES_8dx are shown in the same mesh. (a) C. Meneveau, “Statistics of turbulence subgrid-scale stresses: Necessary
Filtered DNS; (b) SMAG_16dx; (c) DSMAG_16dx. conditions and experimental tests,” Phys. Fluids 6, 815 (1994).
4
U. Piomelli, Y. Yunfang, and R. Adrian, “Subgrid-scale energy transfer
and near-wall turbulence structure,” Phys. Fluids 8, 215 (1996).
5
J. O’Neil and C. Meneveau, “Subgrid-scale stresses and their modelling in
a turbulent plane wake,” J. Fluid Mech. 349, 253 (1997).
= 2⌬x, one option, often used in practice, corresponds to in- 6
C. B. da Silva and O. Métais, “On the influence of coherent structures
terpolating (linearly) values of ␾ into two fictitious nodes upon interscale interactions in turbulent plane jets,” J. Fluid Mech. 473,
placed at ␾i−1/2 and ␾i+1/2. After this, a mean value of ␾i−1/2 7
103 (2002).
J. Silvestrini, “Simulation des grandes échelles des zones de mélange;
and ␾i+1/2 is used in the integral, which yields application à la propulsion solide des lanceurs spatiaux,” Ph.D. thesis,

冕 +⌬x INPG, Grenoble, 1996.


¯ 2⌬x = 1
8
B. Vreman, B. Geurts, and H. Kuerten, “Large eddy simulation of the
␾ i ␾共x⬘兲dx⬘ turbulent mixing layer,” J. Fluid Mech. 339, 357 (1997).
2⌬x −⌬x

冋 册
9
J. A. Domaradzki and E. M. Saiki, “A subgrid-scale model based on the
1 共␾i+1 + ␾i兲 共␾i + ␾i−1兲 estimation of unresolved scales of turbulence,” Phys. Fluids 9, 2148
= ⌬x + ⌬x (1997).
2⌬x 2 2 10
J. A. Domaradzki and K. C. Loh, “The subgrid-scale estimation model in
the physical space representation,” Phys. Fluids 11, 2330 (1999).
␾i+1 ␾i ␾i−1 11
J. A. Domaradzki, K. C. Loh, and P. P. Yee, “Large eddy simulations using
= + + . 共A3兲 the subgrid-scale estimation model and truncated Navier–Stokes dynam-
4 2 4
ics,” Theor. Comput. Fluid Dyn. 15, 421 (2001).
12
B. Geurts, “Inverse modeling for large-eddy simulation,” Phys. Fluids 9,
The problem with this approach is that the interpolating op-
3585 (1997).
eration involves another implicit filtering operation. There- 13
G. S. Winckelmans, A. A. Wray, O. V. Vasilyev, and H. Jeanmart,
fore this method is not equivalent to a true box filter. The “Explicit-filtering large-eddy simulation using the tensor-diffusivity model

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
4534 Phys. Fluids, Vol. 16, No. 12, December 2004 C. B. da Silva and J. C. F. Pereira

33
supplemented by a dynamic Smagorinsky term,” Phys. Fluids 13, 1385 C. Speziale, “Galilean invariance of subgrid-scale stress models in the
(2001). large-eddy simulation of turbulence,” J. Fluid Mech. 156, 55 (1985).
14 34
S. Stolz and N. Adams, “An approximate deconvolution procedure large- C. Canuto, M. Y. Hussaini, A. Quarteroni, and T. A. Zang, Spectral Meth-
eddy simulation,” Phys. Fluids 11, 1966 (1999). ods in Fluid Dynamics (Springer, Berlin, 1987).
15 35
S. Stolz, N. Adams, and L. Kleiser, “An approximate deconvolution J. H. Williamson, “Low-storage Runge–Kutta schemes,” J. Comput. Phys.
method for large-eddy simulation with application to incompressible wall- 35, 48 (1980).
36
bounded flows,” Phys. Fluids 13, 997 (2001). S. Stanley and S. Sarkar, “Influence of nozzle conditions and discrete
16
C. Brun and R. Friedrich, “The spatial velocity increment as a tool for sgs forcing on turbulent planar jets,” AIAA J. 38, 1615 (2000).
37
modeling,” in Modern Simulation Strategies for Complex Turbulent S. Stanley, S. Sarkar, and J. P. Mellado, “A study of the flowfield evolution
Flows, edited by B. Geurts (Edwards, Flourtown, PA, 2001), p. 57. and mixing in a planar turbulent jet using direct numerical simulation,” J.
17
P. Chassaing, Turbulence en Mécanique des Fluides (Cépaduès-Éditions, Fluid Mech. 450, 377 (2002).
38
Toulouse, France, 2000), p. 110. S. Liu, C. Meneveau, and J. Katz, “On the properties of similarity subgrid-
18
J. R. Mansfield, O. M. Knio, and C. Meneveau, “A dynamic les scheme scale models as deduced from measurements in a turbulent jet,” J. Fluid
for vorticity transport equation: Formulation and a priori tests,” J. Com- Mech. 275, 83 (1994).
39
put. Phys. 145, 693 (1998). R. Rogallo and P. Moin, “Numerical simulation of turbulent flows,” Annu.
19
J. Smagorisky, “General circulation experiments with the primitive equa- Rev. Fluid Mech. 16, 99 (1984).
40
tions,” Mon. Weather Rev. 91, 99 (1963). J. Weiss, “The dynamics of enstrophy transfer in two-dimensional hydro-
20
L. Shao, S. Sarkar, and C. Pantano, “On the relationship between the mean dynamics,” LJI-TN-121, La Jolla Institute, San Diego, CA, 1981.
41
flow and subgrid stresses in large eddy simulations of turbulent shear J. Weiss, “The dynamics of enstrophy transfer in 2-dimensional hydrody-
flows,” Phys. Fluids 11, 1229 (1999). namics,” Physica D 48, 273 (1991).
21 42
R. H. Kraichnan, “Eddy viscosity in two and three dimensions,” J. Atmos. J. C. R. Hunt, A. A. Wray, and P. Moin, Eddies, stream, and convergence
Sci. 33, 1521 (1976). zones in turbulent flows, Annual Research Briefs, Center for Turbulence
22
M. Lesieur, Turbulence in Fluids, 3rd ed. (Kluwer, Dordrecht, 1997). Research, Stanford, 1988.
23 43
O. Métais and M. Lesieur, “Spectral large-eddy simulation of isotropic and E. Gutmark and I. Wygnansky, “The planar turbulent jet,” J. Fluid Mech.
stably stratified turbulence,” J. Fluid Mech. 239, 157 (1992). 73, 465 (1976).
24 44
G. K. Batchelor, The Theory of Homogeneous Turbulence (Cambridge R. Ramparian and M. S. Chandrasekhara, “Lda measurements in plane
University Press, Cambridge, 1953). turbulent jets,” ASME J. Fluids Eng. 107, 264 (1985).
25 45
M. Lesieur and J. P. Chollet, “The decay of kinetic-energy and temperature F. O. Thomas and K. M. K. Prakash, “An investigation of the natural
variance in three-dimensional isotropic turbulence,” J. Fluid Mech. 30, transition of an untuned planar jet,” Phys. Fluids A 3, 90 (1991).
46
1278 (1986). R. Akhavan, A. Ansari, S. Kang, and N. Mangiavacchi, “Subgrid-scale
26
D. C. Leslie and G. L. Quarini, “The application of turbulence theory to interactions in a numerically simulated planar turbulent jet and implica-
the formulation of subgrid modelling procedures,” J. Fluid Mech. 91, 65 tions for modelling,” J. Fluid Mech. 408, 83 (2000).
47
(1979). A. Tsinober, M. Ortenberg, and L. Shtilman, “On depression of nonlinear-
27
F. Ducros, P. Comte, and M. Lesieur, “Large-eddy simulation of transition ity in turbulence,” Phys. Fluids 11, 2291 (1999).
48
to turbulence in a boundary layer developing spatially over a flat plate,” J. B. Galati and A. Tsinober, “Self-amplification of the field of velocity
Fluid Mech. 326, 1 (1996). derivatives in quasiisotropic turbulence,” Phys. Fluids 12, 3097 (2000).
28 49
C. B. da Silva and O. Métais, “Vortex control of bifurcating jets: A nu- S. Liu, J. Katz, and C. Meneveau, “Evolution and modeling of subgrid
merical study,” Phys. Fluids 14, 3798 (2002). scales during rapid straining of turbulence,” J. Fluid Mech. 387, 281
29
M. Germano, U. Piomelli, P. Moin, and W. Cabot, “A dynamic subgrid- (1999).
50
scale eddy viscosity model,” Phys. Fluids A 2, 1760 (1991). C. Hartel and L. Kleiser, “Analysis and modeling of subgrid-scale motions
30
U. Piomelli and J. R. Chasnov, “Large eddy simulations: Theory and in near wall turbulence,” J. Fluid Mech. 356, 327 (1998).
51
applications,” in Turbulence and Transition Modeling, edited by M. Hall- O. Métais and M. Lesieur, “Statistical predictability of decaying turbu-
back, D. S. Henningson, A. V. Johannson, and P. H. Alfredsson (Kluwer, lence,” J. Atmos. Sci. 43, 857 (1986).
52
Dordrecht, 1996). I. Dubief and F. Delcayre, “On coherent-vortex identification in turbu-
31
J. Bardina, J. H. Ferziger, and W. C. Reynolds, “Improved subgrid model lence,” J. Turbul. 1, 011 (2000).
53
for large-eddy simulation,” AIAA Pap. 80–1357 (1980). Even if the contrary is not necessarily true: e.g., regions of high rate of
32
V. de Borue and S. Orszag, “Local energy flux and subgrid-scale statistics strain also have high vorticity but do not necessarily imply the presence of
in three dimensional turbulence,” J. Fluid Mech. 366, 1 (1998). coherent vortices.

Downloaded 29 Jan 2009 to 193.136.129.111. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

You might also like