Sei sulla pagina 1di 9

International Journal of Engineering Science 59 (2012) 6573

Contents lists available at SciVerse ScienceDirect

International Journal of Engineering Science


journal homepage: www.elsevier.com/locate/ijengsci

Geometric evolution of the Reynolds stress tensor


S. Gavrilyuk a,, H. Gouin b
a
b

University of Aix-Marseille & C.N.R.S. U.M.R. 6595, IUSTI, 5 rue E. Fermi, 13453 Marseille Cedex 13, France
University of Aix-Marseille & C.N.R.S. U.M.R. 7340, Case 322, Av. Escadrille Normandie-Niemen, 13397 Marseille Cedex 20, France

a r t i c l e

i n f o

Article history:
Received 14 March 2011
Received in revised form 21 September 2011
Accepted 21 November 2011
Available online 11 April 2012
Dedicated to Professor Victor Berdichevsky
on the occasion of his 65th birthday
Keywords:
Reynolds stress tensor
Capillary uids
Variational principle

a b s t r a c t
The dynamics of the Reynolds stress tensor for turbulent ows is described with an evolution equation coupling both geometric effects and turbulent source terms. The effects of
the mean ow geometry are shown up when the source terms are neglected: the Reynolds
stress tensor is then expressed as the sum of three tensor products of vector elds which
are governed by a distorted gyroscopic equation. Along the mean ow trajectories, the uctuations of velocity are described by differential equations whose coefcients depend only
on the mean ow deformation. If the mean ow vorticity is small enough, an approximate
turbulence model is derived, and its application to shear shallow water ows is proposed.
Moreover, the approximate turbulence model admits a variational formulation which is
similar to the one of capillary uids.
2012 Elsevier Ltd. All rights reserved.

1. Introduction
The system describing turbulent compressible barotropic ows is composed of equation of mass balance, equation of
average momentum and evolution equation for the Reynolds stress tensor. In the following, we see that the Reynolds stress
tensor equation is mainly driven by the velocity gradient tensor of the mean motion and this equation is the main object of
our study when the source term is negligible.
The reason for considering the simplied turbulence model without the source term is twofold. First, in numerical studies
of compressible turbulent ows, this homogeneous equation is a natural step in applying the splitting-up technique (see for
example (Berthon, Coquel, Hrard, & Uhlmann, 2002)). Secondly, such a homogeneous system appears as an exact asymptotic model of weakly shearing ows of long waves over a at bottom (see (Teshukov, 2007)). The only difference is the space
dimension: two dimensions are considered for shallow water ows instead of three dimensions for the general case.
We use the spectral decomposition of the Reynolds stress tensor and in the homogeneous case, we obtain a simpler
dynamical system for the eigenvalues and the eigenvectors of the Reynolds stress tensor. The system admits a simple physical interpretation: the motion of each point of the turbulent ow is analogous to the motion of a free rigid body moving
along the mean ow and rotating with an angular velocity which is different from the mean ow vorticity. The angular velocity is completely determined by the mean ow velocity. The moments of inertia of the free rigid body are not constant, they
are also determined by the mean ow.
When the mean ow vorticity is small enough, an approximate turbulence model is obtained, which admits a variational
formulation.

Corresponding author.
E-mail addresses: sergey.gavrilyuk@polytech.univ-mrs.fr (S. Gavrilyuk), henri.gouin@univ-amu.fr (H. Gouin).
0020-7225/$ - see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijengsci.2012.03.008

66

S. Gavrilyuk, H. Gouin / International Journal of Engineering Science 59 (2012) 6573

2. The governing equations


The governing equations of barotropic turbulent compressible uids are (see (Mohammadi & Pironneau, 1994; Pope,
2005; Wilcox, 1998)):

8
>
< hqit hqiU i ;i 0;
hqiU i t hqiU i U j hpidij hqui uj i;j 0;
>
: hqu u i hqu u iU hqu u iU hqu u iU S
i j t
i j
i k
ij
k ;k
k j
i;k
j;k

where brackets mean the averaging, coma means the derivation with respect to the Eulerian coordinates x = {xi},
i 2 {1, 2, 3} and index t means the partial derivative with respect to time, q is the uid density, U = {Ui}, i 2 {1, 2, 3} is the
mass average velocity, p is the pressure, u = {ui}, i 2 {1, 2, 3} is the velocity uctuation verifying hqui = 0. Repeated indices
mean summation. Here S = {Sij} is a source term, and its explicit expression can be written as:

Sij hui p;j i  huj p;i i  hqui uj uk i;k :


We introduce the Reynolds stress tensor

R hqu  ui;

Rij hqui uj i:

System (1) can be rewritten in the tensorial form

8
@hqi
>
>
divhqiU 0;
>
>
>
@t
>
>
<
dU
hqi
rhpi div RT 0;
dt
>
>
>
 T
> dR
>
@U
@U
>
>
S;
R divU
RR
:
dt
@x
@x

where d/dt means the material derivative with respect to the mean motion

d
@
UT r:
dt @t
The superscript T means the transposition. Using the mass conservation law, the equation for the volume Reynolds stress
tensor R can be rewritten as the equation for the specic (or per unit mass) Reynolds stress tensor

 T
dP @U
@U
S

PP

;
dt @x
@x
hqi

where

R
:
hqi

The structure of source term S has generated much debate within physical and mathematical communities. Our goal is not to
add new closure hypotheses, but to study the structure of the master Eq. (3) when S = 0. The reason is twofold:
rstly, in the numerical study of compressible turbulent ows, this is a natural step in applying the splitting-up technique
(see for example (Berthon et al., 2002)),
secondly, system (2) also appears as an exact asymptotic model of weakly shearing ows of long waves (turbulent shallow water ows) over a at bottom (see (Teshukov, 2007)):

8
@h
>
>
divhU 0;
>
>
@t
>
>
!
>
>
2
< dU
gh
div RT 0;
r
h
dt
2
>
>
>
>
 T
>
>
>
dR
@U
@U
>
:
R divU
RR
0:
dt
@x
@x

In system (4), h is the uid depth playing the role of the average density; the average pressure is given by hpi = gh2/2, g is
the gravity acceleration, and

Z
0

h

 
i
e U  U
e  U dz;
U

hU

Z
0

e
Udz;

67

S. Gavrilyuk, H. Gouin / International Journal of Engineering Science 59 (2012) 6573

e is the instantaneous velocity. Equations are written for three-dimensional long waves and the production term is
where U
zero in the limit of weakly shearing ows. Eq. (4) are hyperbolic (see (Teshukov, 2007) for proof). Finally, we will focus on the
equation of the Reynolds stress tensor per unit mass

 T
dP @U
@U

PP
0:
dt @x
@x

The case S = 0 corresponds to conservative motions of turbulent compressible ows; these motions verify the equations

8
@hqi
>
>
divhqiU 0;
>
>
>
@t
>
>
>
<
dU
rhpi div RT 0;
hqi
dt
>
>
>
 T
>
>
>
dR
@U
@U
>
>

R
divU

R
0:
:
dt
@x
@x

The particular case rotU = 0 was investigated in Debive, Gouin, and Gaviglio (1982). In such a case (@U/@x)T = @U/@x and Eq.
(5) corresponds to a two-covariant tensor convected by the mean ow. This means that P has a zero Lie derivative dL with
respect to the velocity eld U and the tensor P0, image of P in Lagrange coordinates (t, X), only depends on X = {Xi}, i 2 {1, 2, 3}

dL P 

dP @U
@U

PP
0;
dt @x
@x

P F T 1 P0 XF 1 ;

where F = @x/@X is the deformation gradient of the mean motion.


The aim of the paper is the study of the homogeneous Reynolds stress tensor equation structure (5) in the case rotU 0.
3. Geometric properties of the Reynolds stress tensor evolution
The Reynolds stress tensor P is symmetric and semi-positive denite. The tensor P can be rewritten in a local basis of
orthonormal eigenvectors in the form

3
P

k2a ea  ea 

a1

3
P

a1

k2a ea eTa :

The eigenvalues k2a ; a 2 f1; 2; 3g are non-negative; the case k2a > 0 is a generic one. For the two-dimensional case, k23  0. Let
us denote

aa ka ea ;

ka > 0 a 2 f1; 2; 3g:

Then,

3
P

3
P

aa  aa 

a1

aa aTa :

a1

From Eq. (8), we deduce

"

T #
3
dP P
daa T
daa
:

aa aa
dt a1 dt
dt

By using Eq. (9), Eq. (5) can be written




 T
3
P
daa @U
daa @U
aa aTa
aa aTa 0:

@x
@x
dt
a1 dt

10

The vector daa/dt + (@U/@x)aa can be developed in the local basis {ab}, b 2 {1, 2, 3} of eigenvectors; one obtains
3
P
daa @U
aa Aba ab ;

@x
dt
b1

a 2 f1; 2; 3g

11

where Aba, (a, b 2 {1, 2, 3}) are the scalar components to be determined. By using Eq. (11), Eq. (10) leads to
3
P

Aaa gaa

a1

3
P

Aab Aba gab 0;

ab1

where gaa 2aa aTa and gab gba aa aTb ab aTa ; a; b 2 f1; 2; 3g, are six independent symmetric tensors. Consequently,

Aaa 0 and Aab Aba 0;

a; b 2 f1; 2; 3g:

68

S. Gavrilyuk, H. Gouin / International Journal of Engineering Science 59 (2012) 6573

Eq. (10) is equivalent to

daa @U
aa K  ipea

@x
dt

with p A32 e1 A13 e2 A21 e3 ;

a 2 f1; 2; 3g;

12

where a diagonal matrix K and an antisymmetric matrix i(p) are determined in the basis eb, b 2 {1, 2, 3} as

k1

KB
@0
0

k2
0

C
0 A;
k3

B
ip @ A21
A13

A13

A21

C
A32 A:

0
A32

The vectors ab, b 2 {1, 2, 3} are orthogonal, aTa ab 0; a b. This is equivalent to

aTa

dab
daa
aTb
0:
dt
dt

13

Consequently, Eqs. (12), (13) yield

8a b 2 f1; 2; 3g;





@U
@U
ab aTb K  ipea 
aa 0;
aTa K  ipeb 
@x
@x

or

2ka kb eTa Deb aTa K  ipeb aTb K  ipea  k2a eTa ipeb k2b eTb ipea ;

14

where

 T !
1 @U
@U

D
2 @x
@x
is the rate of deformation tensor corresponding to the mean ow. We denote the mixed product of three vectors {a, b, c} as
(a, b, c)  aT(b ^ c). Hence, Eq. (14) yields







2ka kb eTa Deb k2a  k2b ea ; p; eb k2b  k2a p; ea ; eb k2b  k2a pT ec ;
where {a, b, c} is a cyclic permutation of the triplet {1, 2, 3}. Finally, we get

pT ec

2ka kb eTa Deb


k2b  k2a

15

Eq. (12) can be written

ka

dea dka
@U
ea  K  ipea 0;

ea ka
@x
dt
dt

eTa

dea
0;
dt

a 2 f1; 2; 3g:

16

Since

eTa K  ipea ka eTa ipea 0;

by multiplying the left side of Eq. (16) with eTa , we get


dka  T
ea Dea ka 0:
dt



By multiplying the left side of Eq. (16) with the projector I  ea eTa , we get

ka


@U
dea
ea  K  ipea 0:
ka I  ea eTa
@x
dt

We will prove that there exists a vector P such that

dea
P ^ ea :
dt
Such a vector P should verify the condition


@U
ea  K  ipea 0:
ka iPea ka I  ea eTa
@x

17

S. Gavrilyuk, H. Gouin / International Journal of Engineering Science 59 (2012) 6573

69

By multiplying Eq. (17) with eTb where b a, we get

ka PT ec kb pT ec  ka eTb

@U
ea :
@x

18

By replacing Rel. (15) into Eq. (18) we get

k2b eTa
PT e c

 T !
@U
@U
eb

@x
@x
k2b

 ka

 eTb

@U
ea
@x

eTb k2b

!
 T
@U
@U
ea
k2a
@x
@x
k2b  k2a

Now, we can formulate the result,


Theorem 1. The Reynolds stress tensor can be written in the form

R hqi

3
P

a1

k2a ea  ea :

The eigenvectors ea and the eigenvalues ka verify the equations:

8
dea
>
>
P ^ ea ;
<
dt


2
>
>
: d Lnka 2l ;
a
dt

a 2 f1; 2; 3g;

19

where

eTb

la eTa Dea ; PT ec

k2b

!
 T
@U
2 @U
ea
ka
@x
@x
k2b  k2a

The triplet {a, b, c} corresponds to a cyclic permutation of the triplet {1, 2, 3}.
Eq. (191) is similar to the equations of a rigid body (see (Marsden & Ratiu, 1994)). The vectors ea form a natural moving
frame fea g3a1 whose evolution is determined by the mean rate of deformation tensor. The eigenvalues k2a of the Reynolds
stress tensor are determined by the evolution Eq. (192). Let us note that, if ka are initially positive, they will be positive
for any time. Hence, it means that the tensor P will always be positive denite. Due to the mass conservation law (21)
and Eq. (192), we obtain the following quantity conserved along trajectories of mean ow:



3
Q
d
hqi2 k2a 0:
dt
a1
Consequently, system (19) admits an invariant scalar along the trajectories of mean ow. This invariant was earlier obtained
in Debive, Gouin, and Gaviglio (1982); Debive et al. (1982) in a different form. Let us introduce the turbulent specic
energy

eT

3
1
1P
tr P
k2 :
2
2 a1 a

In the incompressible (isochoric) case, we have dhq i/dt = 0; the turbulent energy is minimal in the isotropic case when the


three eigenvalues k2a are equal k21 k22 k23 k2 . In this case, the orthonormal eigenvectors ea, a 2 {1, 2, 3} of the Reynolds
stress tensor P are also the orthonormal eigenvectors of the mean rate of deformation tensor D and la are the corresponding
eigenvalues.
In the compressible isotropic case eT = hqi2/3j, j = 3k2/(2h qi2/3), and j is a classical invariant of isotropic compressible
turbulence.
Q
In presence of shock waves the quantity hqi2 3a1 k2a is not conserved through shocks; it increases like the classical entropy in compressible uid dynamics. The estimation of the jump of turbulence entropy in isotropic case was given in Gavrilyuk and Saurel (2006).
As a consequence, the governing Eq. (6) admit the energy conservation law

 





@
1
1
div hqiU jUj2 ei eT hpiI RU 0
hqi jUj2 ei eT
@t
2
2
where the internal specic energy ei is dened by



1
dei hpid qi
h

70

S. Gavrilyuk, H. Gouin / International Journal of Engineering Science 59 (2012) 6573

and the mean pressure hpi is supposed to be a given function of hqi. Indeed, using (192) we immediately obtain

 





@
1
1
deT
div hqiU jUj2 ei eT hpiI RU hqi
hqi jUj2 ei eT
trRD
@t
2
2
dt
 3 
 3

P 2
hqi d P 2
k hqitr
ka la 0:

2 dt a1 a
a1
System (6) is a conservative and Galilean invariant system of equations which is the counterpart of the Euler equations for
the turbulent compressible ows. The equations for the Reynolds stress tensor (63) are rewritten in a simpler form admitting
a clear physical interpretation.
4. An approximate model
In this Section, we derive a useful approximation of model (6) describing compressible turbulent ows for motions characterized by a small average vorticity.
4.1. Preliminaries
Eq. (5) can be rewritten as:

 T
dP
@U
@U
PP

irotU; P
dt
@x
@x

20

with

irotU; P P irotU  irotUP;

irotU

 T
@U
@U

:
@x
@x

Let s be a characteristic time scale and x a characteristic value of the mean vorticity norm krot Uk; we assume that

sx  1:

21

Relation (21) is veried, in particular, for motions which are close to one-dimensional ones. Eq. (20) gets the form

 T
dP
@U
@U

0:
PP
dt
@x
@x

22

Using the solution (7) when P0(X) = I, we consider the Reynolds stress tensor in the form

P
a

rua  rua ;

where index a ranges over a nite number of integers and ua are generalized Lagrangian coordinates:

dua
0:
dt
Covectors
T

ba

@ ua
@x

verify the identity


T

dba
T @U
0;
ba
@x
dt
T

corresponding to a zero Lie derivative of ba with respect to the mass average velocity. In such a case, Eq. (22) is identically
veried. Symmetric tensor P is determined by six scalar elds ua, a 2 {1, . . . , 6}. If, initially, vectors rua, a 2 {1, 2, 3} constitute an orthogonal system corresponding to the eigenvectors of P, we can choose ua = 0, a 2 {4, 5, 6} and consequently,
3
P

a1

rua  rua :

23

It is worth to note that since the Reynolds stress tensor is positive denite, it can be considered as a metric tensor of a Riemannian space associated with the metric

Pij

3 @u @u
P
a
a
:
a1 @xi @xj

S. Gavrilyuk, H. Gouin / International Journal of Engineering Science 59 (2012) 6573

71

This metric is at because


3
P
P
Pij dxi dxj
dua 2

a1

i;j

It is interesting to note that this special structure of the Reynolds stress tensor implies a variational structure of Eq. (6).
4.2. The Hamilton principle
The aim of this Section is to prove that in special form (23), System (6) admits a variational formulation which is similar to
the one of capillary uids (Cahn & Hilliard, 1958; Casal, 1963, 1972; DellIsola, Gouin, & Rotoli, 1996; Eglit, 1965; Truskinovsky, 1982; van der Waals, 1979). However, in our case, the expression of the capillary energy is determined by the gradients of
three scalar order parameters transported along the trajectories of the mean ow.
We consider the specic internal energy in the form

ei ei hqi:
The mean density is submitted to the constraint

@hqi
divhqiU 0:
@t
Let us dene the specic turbulent energy as

eT

3 jru j2
P
a
;
2
a1

where scalars ua are submitted to the constraint

dua
0;
dt

a 2 1; 2; 3:

For a material volume Dt of the mean motion, the Hamilton action calculated between times t1, t2 is

t2

Z Z Z

t1


hqiLdv dt;

Dt

with the specic Lagrangian



1 T
U U  ei  eT :
2

The uid motion is a C2-diffeomorphisme / from a three-dimensional space D0 into the physical space Dt:

x /X; t or xi /i X 1 ; X 2 ; X 3 ; t;

i 2 f1; 2; 3g:

Let a one-parameter family of virtual motions denoted by {/e}, possessing continuous derivatives up to the second order and
expressed in the form:

x UX; t; e;
with e 2 O; where O is an open interval containing 0 and such that U(X, t; 0) = /(X, t) (the real motion of the continuous medium is obtained when e = 0). The derivation with respect to e when e = 0 is denoted by d. Derivation d is named variation and
the virtual displacement d/ is the variation of the motion of the medium. At time t, the virtual displacement of the particle x
is dx obtained when dX = 0 and de = 1 at e = 0; the virtual displacement corresponds to the eld of tangent vectors to Dt

x 2 Dt ! f wx 

@U

2 T x Dt ;
@ e
e0

where Tx(Dt) is the tangent vector bundle to Dt at x.


The Hamilton principle reads: for each vector eld of virtual displacements such that f and its derivatives vanish at the boundary @ X of X,

da 0:
We have the following general results (see (Berdichevsky, 2009; Casal, 1963; Gavrilyuk & Gouin, 1999; Serrin, 1959)):

  R
8 R t RR R
2
>
d
h
q
iLd
v
dt X hqidLdv dt;
>
t
D
t
1
>
>
>
>
>
df
>
<
dU ;
dt
>
>
dhqi hqidivf;
>
>


>
>
@ ua @f
> @ ua
>
:d

;
@x
@x @x

72

S. Gavrilyuk, H. Gouin / International Journal of Engineering Science 59 (2012) 6573

where X = [t1,t2]  Dt and

hqidL hqiUT

R
X

is the quadruple integral in the timespace domain X. Consequently,


T
3
P
df
@ei
@ ua @f @ ua
:
hqi2
divf hqi
dt
@hqi
@x @x @x
a1

Let us denote hpi = hqi2@ei/@hqi, the mean pressure scalar eld of the uid; due to the identities,

8
>
@hqiUT f
dUT
T df
T
>
>


divh
h
q
iU
q
iU
f

h
q
i
f;
>
>
dt
@t
dt
>
>
>
<
@hpi
f;
hpidivf  divhpif 
@x
>
!
!
!
>



T

T

T
>
T
>P
3
3
3
3
>
P
P
P
@ ua @f @ ua
@ ua @ ua @f
@ ua @ ua
@ ua @ ua
>
>

h
q
i

tr
h
q
i
div
h
q
i
div
h
q
i
f

f:
>
:
@x @x @x
@x
@x @x
@x
@x
@x
@x
a1
a1
a1
a1
!

T
@hqiUT f
@ ua @ ua
; divhqiUT f; divhpif and div hqi
f can be integrated on @ X
Stokes formula implies that terms
@t
@x
@x
where f vanishes. We get for each eld of virtual motion,

!)

T
Z (
3
dUT @hpi P
@ ua @ ua
da 
hqi
div hqi

f dv dt;
@x
dt
@x
@x
X
a1
and the fundamental lemma of variational calculus yields the equation of motion

hqi



3
P
dU
rhpi div hqi rua  rua 0;
dt
a1

the variational formulation of the approximate system is established.


5. Conclusion
The equations of uid turbulent motions take three equations into account: the equation of mass balance (11), the balance
equation of average momentum (12), and the Reynolds stress tensor equation of evolution (13); this last equation has been
the object of our study. If the turbulent sources are neglected, the turbulent uid motion is a superposition of the mean motion and turbulent uctuations. The eigenvectors of the Reynolds stress tensor carry the uctuations associated with the
mean ow deformation. The amplitudes of turbulent deformations are dened by the eigenvalues of the Reynolds stress tensor. The equations for the directions of turbulent uctuations are reminiscent of a gyroscopic type equation for the motion of
a free rigid body (Eq. (191)). The amplitude evolution of turbulent deformations are determined by the diagonal values la of
the mean rate of deformation tensor D expressed in the eigenvector basis of the Reynolds stress tensor. The turbulence increases with the time when la < 0, and decreases when la > 0. In the particular case of incompressible uid motions we have
tr D = 0, and hence there always exists a direction in which the turbulence is increasing while in other directions the turbulence is decreasing. Over the past two decades great progress has been made in understanding many aspects of the kinematics and dynamics of a wide variety of turbulent ows as a result of access to the mean velocity gradient tensor (see
(Ganapathisubramani, Longmire, & Marusic, 2006; Wallace, 2009)). Such an access could be used for an experimental determining of these eigenvalues.
In the case of a small mean vorticity, a new approximate model admitting a variational formulation is derived.
Acknowledgement
We thank Professor Vladimir Teshukov who drew our attention to the model of weakly sheared ows during his visit in
Marseille in the fall of year 2007, and Professor Boris Kolev for valuable discussions.
References
Berdichevsky, V. (2009). Variational principles of continuum mechanics, I. Fundamentals. Berlin: Springer.
Berthon, C., Coquel, F., Hrard, J. M., & Uhlmann, M. (2002). An approximate solution of the Riemann problem for a realisable second-moment turbulent
closure. Shock Waves, 11, 245269.
Cahn, J. W., & Hilliard, J. E. (1958). Free energy of a nonuniform system. I. Interfacial free energy. Journal of Chemical Physics, 28, 258267.
Casal, P. (1963). Capillarit interne en mcanique des milieux continus. C.R. Academic Science Paris, 256, 38203822.
Casal, P. (1972). La thorie du second gradient et la capillarit. C.R. Academic Science Paris, 274, 15711574.
Debive, J.-F., Gouin, H., & Gaviglio, J. (1982). Momentum and temperature uxes in a shock waveturbulence interaction. In Z. Zaric (Ed.), Proceedings of the
ICHMT/IUTAM symposium on the structure of turbulence and heat and mass transfer (pp. 277296). London: Hemisphere Publishing Corporation.
Debive, J.-F., Gouin, H., & Gaviglio, J. (1982). Evolution of the Reynolds stress tensor in a shock waveturbulence interaction. Indian Journal of Technology, 20,
9097.
DellIsola, F., Gouin, H., & Rotoli, G. (1996). Nucleation of spherical shell-like interfaces by second gradient theory: Numerical simulations. European Journal
of Mechanics B/Fluids, 15, 545568.
Eglit, M. E. (1965). A generalization of the model of an ideal compressible uid. Journal of Applied Mathematics and Mechanics, 29, 395399.

S. Gavrilyuk, H. Gouin / International Journal of Engineering Science 59 (2012) 6573

73

Ganapathisubramani, B., Longmire, E. K., & Marusic, I. (2006). Experimental investigation of vortex properties in a turbulent boundary layer. Physics of Fluids,
18, 055105.
Gavrilyuk, S., & Gouin, H. (1999). A new form of governing equations of uids arising from Hamiltons principle. International Journal of Engineering Science,
37, 14851520.
Gavrilyuk, S., & Saurel, R. (2006). Estimation of the turbulence energy production across a shock wave. The Journal of Fluid Mechanics, 549, 131139.
Marsden, J. E., & Ratiu, T. S. (1994). Introduction to mechanics and symmetry. Series in applied mathematics (vol. 17). Berlin: Springer.
Mohammadi, B., & Pironneau, O. (1994). Analysis of the K-epsilon turbulence model. Research in applied mathematics. New York: John Wiley & Sons.
Pope, S. B. (2005). Turbulent ows. Cambridge University Press.
Serrin, J. (1959). Mathematical principles of classical uid mechanics. In Encyclopedia of Physics VIII/1 (pp. 125263). Berlin: Springer.
Teshukov, V. M. (2007). Gas dynamic analogy for vortex free-boundary ows. Journal of Applied Mechanics and Technical Physics, 48, 303309.
Truskinovsky, L. (1982). Equilibrium phase boundaries. Soviet Physics Doklady, 27, 551553.
van der Waals, J. D. (1979). The thermodynamic theory of capillarity under the hypothesis of a continuous density variation. J. Stat. Phys., 20, 197244.
Translated by J. S. Rowlinson (1893).
Wallace, J. M. (2009). Twenty years of experimental and direct numerical simulation access to the velocity gradient tensor: What have we learned about
turbulence? Physics of Fluids, 21, 021301.
Wilcox, D. C. (1998). Turbulence modeling for CFD. La Caada, California: DCW Industries, Inc.

Potrebbero piacerti anche