You are on page 1of 34

Journal of Scientific Computing, Vol. 7, No.

1, 1992

Energy and Spectral Dynamics in


Decaying Compressible Turbulence
Shigeo Kida 1 and Steven A. Orszag 2

Received August 6, 1991

The statistical properties of decaying compressible turbulence are investigated


by direct numerical simulations of flow in a periodic cube. Starting with fully
developed turbulence for various microscale Reynolds numbers R~, rms Mach
numbers M, and small- and large-scale compressive ratios rcs and rCL, we find
that the kinetic energy exhibits an exponential decay in time. Interactions
between the compressive and rotational components of kinetic energy are
weaker than self-interactions of the respective components. The energy spectrum
of rotational component obeys the same Kolmogorov similarity law as for
incompressible turbulence and forced compressible turbulence. The form of the
energy spectrum of the compressive component, on the other hand, depends
strongly on M. As the turbulence decays freely, R~. and M decrease in time but
rcs and rCL tend to some universal values.

KEY WORDS: Compressible turbulence; numerical simulation; spectral


dynamics.

1, INTRODUCTION
The effects of compressibility become increasingly significant in the
dynamics of compressible turbulent flows as the fluctuating (rms) Mach
number M increases. When the fluctuating turbulent velocity becomes
comparable with the speed of sound, interactions between vorticity and
sound waves are substantial. The most relevant parameter characterizing
incompressible turbulence is the microscale Reynolds number R~. The
statistical properties of turbulence, such as the form of the energy spec-
trum, the skewness and flatness of velocity-derivatives, for a given R~. are
more or less universal irrespective of forcing or boundary conditions.

J Research Institute for Mathematical Sciences, Kyoto University, Kyoto 606, Japan.
2 Applied and Computational Mathematics, Princeton University, Princeton, NJ 08544.
1

854/7/1-1 0885-7474/92/0100-0001506.50/0 9 1992 Plenum Publishing Corporation


2 Kida and Orszag

For compressible turbulence, there are additional parameters that


characterize the flow. In a previous paper (Kida and Orszag, 1990a; here-
after this paper will be referred to as KO) we reported numerical studies
of forced compressible turbulence and showed that the statistics of the
velocity field depend not only on the rms Maeh number but also on the
intensity of fluctuations of thermodynamic variables (e.g., density fluctua-
tion 6p'/(p)) and the small- and large-scale compressive ratios rcs and
rcL. Since the local Mach number may take much larger values than (typi-
cally twice as large as) the rms Mach number, the formation of shock
waves is expected even at moderate rms Mach numbers (~0.5). In this
case, interactions between vorticity field and shock waves are important
(Kida and Orszag, 1990b).
In KO, we decomposed the velocity field into the compressive and
rotational components and examined energy transfer among these two
components and the internal energy through various nonlinear interac-
tions. The coupling between the compressive and rotational components is
relatively weak compared with self-interactions of the respective com-
ponents. The spectrum of the rotational component of energy does not
depend on M and obeys the same Kolmogorov similarity law as for the
incompressible case. The compressive component, on the other hand,
depends strongly on M; in particular, the spectrum decays less rapidly
with wave number for larger M. We observe a periodic energy exchange
between the internal energy and the compressive component of the kinetic
energy for high compressive ratios. This energy exchange occurs through
the pressure-dilatation interaction term. A similar observation had been
made by Passot and Pouquet (1987) in their numerical simulation of two-
dimensional decaying turbulence. The contribution to the energy dissipa-
tion by the compressive component of velocity (longitudinal mode) is
effective at high Mach number and high compressive ratios (see also
Sarkar et al., 1991, and Zeman, 1990).
Previously, we also studied the role of shock waves in generating vor-
ticity (Kida and Orszag, 1990b). We,found that vorticity production due to
baroclinic generation is effective by curved shocks and that the pressure and
density gradients are nearly parallel. These three-dimensional results are
different from the two-dimensional results of Passot and Pouquet (1987).
They observed that vorticity is produced by shock wave collisions near
which pressure and density gradients are nearly orthogonal. The effects of
compressibility at low Mach number were analyzed by Sarkar et al. (1991)
and Erlebacher etal. (1990). They found that different asymptotic states
exist which depend on Mach number and pressure fluctuations of the initial
velocity field.
In this article, we study the statistical properties of freely decaying
Decaying Compressible Turbulence 3

compressible turbulence. We investigate the decay law of kinetic energy,


the Kolmogorov similarity law of the kinetic energy spectrum, the relaxa-
tion of the compressive ratios to some universal values, and the energy
exchange among the internal energy and the compressive and rotational
components of kinetic energy and so on. We also compare these quantities
with those for forced turbulence (see KO). In Sec. 2, we describe the
governing equation and the initial condition used in our numerical simula-
tions. Iri Sec. 3, we study the behavior of various integral quantities of
decaying turbulence such as the energy decay law and energy exchange
among various components of the flow. In Sec. 4, the spectral charac-
teristics of energy, including those of the compressive and rotational parts
of the flow, are compared with the Kolmogorov similarity law. In Sec. 5,
we summarize our conclusions.

2. DYNAMICAL EQUATIONS AND INITIAL CONDITIONS


The motion of a compressible viscous fluid is described by conserva-
tions of mass, momentum, and total energy as

+ ~Xi (P"i)=0 (2.1)

a 0p 2
-(PU)i+~'xj(jpuui)j=~3t -~xi+Yeo~xj (Sij-89 (2.2)

~ 1
VZT
~tEr+-~x~ (ET+p) Ui = 2
Mo pr Reo(?_ 1 )
2 0
+ R%eo~xj ui(Su- 89 (2.3)

where p is the density, p the pressure, T the temperature, u = (ul, u=, u3)
the velocity,

A _=V. u (2.4)

the dilatation, and

1/Ouj Oug) (2.5)

the strain-rate tensor. All the field variables are nondimensionalized in


terms of a reference density Po, velocity Uo, length/o, and temperature To.
The bulk viscosity is assumed to be zero. Terms with repeated indices
4 Kida and Orszag

are summed over those indices. These equations are supplemented by the
equation of state (the perfect-gas law)

pT
p - 7Mo2 (2.6)

The total energy density

ET = Ez + EK (2.7)

is the sum of the internal energy density

E,= p (2.8)
y-1

and the kinetic energy density

E ~ = 2 lul 2 (2.9)

We have three nondimensional parameters: a reference Reynolds


number,

P0 Uolo
Reo - - - (2.10)
kt

a reference Mach number, which is defined by the ratio of a typical velocity


uo and sound speed co= (TRTo) m,

Mo =u~ (2.11)
Co

and the Prandtl number

Pr = Cp# (2.12)
K

where 7 is the specific heat ratio, R the perfect gas constant, Cp the
constant-pressure specific heat,/~ the dynamic viscosity, and x the thermal
conductivity.
We solve (2.1)-(2.3) numerically by a Fourier pseudospectral method
using N 3 (=643) collocation points in a periodic cube of side 2n. Time
stepping is performed by a Runge-Kutta-Gill scheme. The initial flow field
is taken from statistically quasiequilibrium states which were realized by
Decaying Compressible Turbulence 5

introducing a random external force (see KO). Three cases are examined.
Some of the characteristic parameters of the velocity fields at the initial and
final times of the simulation are shown in Table I. Here,
/ 5 \1/2
R;=~-~e) (0)(I.I 2) (2.13)

e being the energy-dissipation rate [-see (3.6) below] and ( ) denoting the
spatial average, is the microscale Reynolds number,

M=((~-)2) 1/2 (2.14)

c = (?RT) 1/2 being sound speed, is the rms Mach number,


(zl ~)
rcs = (A 2) + (1~o12)
(2.15)

is the small-scale compressive ratio, and

(lucl 2)
r~L- (lul2) (2.16)

Uc being the compressive component of velocity, is the large-scale com-


pressive ratio.
The characteristic features of turbulence at the initial time for these
three cases are summarized as follows. For case I, the microscale Reynolds
number is relatively high (Rj~40), the rms Mach number is moderate
(M~0.3), and the vortical motion overwhelms the compressive motion

9 Table I. Characteristic Parameters of Velocity Field

Case R~ M rcs rCL ([ul 2) (1~[2)

At the initial time


I 44 0.33 0.012 0.012 0.087 0.25
II 36 0.80 0.29 0.066 1.1 2.5
III 5.4 0.11 0.998 0.96 9.0 x 10 3 5.6 x 10 -5

At the final time


I 12 0.11 0.017 0.029 0.009 0.032
II 17 0.35 0.027 0.027 0.25 0.90
III 4.3 0.05 0.997 0.74 2.2 x 10 .3 1.1 x 10 -5
6 Kida and Orszag

(rcs ~ rcL ~ 0.01). For case II, the microscale Reynolds number is com-
parable with that for case I, but the rms Mach number~is higher ( M ~ 0 . 8 )
and the compressive motion is not negligible at small scales (rcs ~ 0.3). In
case III, both the microscale Reynolds number and the rms Mach number
are small ( R ~ 5 , M ~ 0 . 1 ) , and in contrast to casesI andII, the com-
pressive motion overwhelms the vortical motion (rcs ~ rcL ~ 1 ). The inten-
sity of density fluctuations ((p - ( p ) ) 2 ) m / ( p ) is another characteristic
parameter, which is 0.05, 0.28, and 0.18 for cases I, II, and III, respectively
(not shown in Table I).
The other parameters in the simulation are set to be M0 = 1, 7 = 1.4,
and Pr = 0.7. The reference Reynold number Re0 is 200, 50, and 100 for
cases I, II, and III, respectively. Finally, the kinetic energy is distributed
isotropically in the wave-number space in cases I and II, but the energy is
localized near the three orthogonal axes in case III (see Fig. 8 in KO).

3. GLOBAL QUANTITIES

3.1. Characteristics of the Velocity Field


The time development of the velocity ficld varies greatly for the
different cases. We first discuss the behavior of global (integral) quantities
of the flow field.
The time evolutions of the microscale Reynolds number R~ofor cases I,
II, and III arc plotted in Figs. la, lb, and lc, respectively. The Reynolds
number decreascs monotonically in time from 44 to 12 for case I and from
36 to 17 for case II. For case III, on the other hand, thc Reynolds number
decreases while undergoing a regular oscillation with a period of 4.1 time
units. As will be seen below, similar oscillations arc also obscrved in other
quantities, such as the rms Mach number, the compressive ratios, the
kinetic and internal energies, etc. These oscillations synchronize with
collisions of sound waves (see KO).
In Figs. 2a, 2b, and 2c we plot the time evolution of the rms Mach
number M and the pointwisr local maximum of the Mach number (]ul/c) 2
for cases I, II, and III, respectively. Note that the local maximum is about
two times larger than the rms Mach number. The rms Mach number
decreases monotonically in time from 0.33 to 0.11 for case I and from 0.80
to 0.35 for case II. The flow is therefore subsonic as a whole for these two
cases. But, as secn in Fig. 2b, the flow can be locally supersonic and shock
waves play a role in dynamics of fluid motion (Kida and Orszag, 1990b)
until t ~ 38. For case III, the Mach number decreases in an oscillating way
as for the Reynolds number.
Decaying Compressible Turbulence 7

~=

t
ii' ioo llo l~o
t
130 1.10 150

"5

Fig. 1. Time evolution of the microscale


Reynolds number R~ for (a) case I, (b) case II,
3o 3'2 3"4 36 3'8 40
t and (c) case III.

It was shown in KO that the characteristics of compressible turbulence


differs greatly depending on the small- and large-scale compressive ratios.
The time development of the small-scale compressive ratio rcs is plotted in
Figs. 3a, 3b, and 3c for cases I, II, and III, respectively. For case I, this
ratio first decreases in time but then increases up to rcs~0.017. For
case II, the ratio decreases very rapidly to rcs ,~ 0.027. These two simula-
tions leads us to the interesting speculation that rcs may relax to a univer-
sal value of around 0.02. For case III, the ratio decreases in time again with
a regular oscillation. Because of the shortness of the simulation period,
however, the compression ratio is still decaying steadily and we cannot say
anything about a possible equilibrium value.
We see similar behavior for the large-scale compressive ratio rCL
in Figs. 4a, 4b, and 4c. Comparing Figs. 4a and 4b, we expect that rcL
8 Kida and Orszag

N
,,,.

.... ' . . . . . . . . . . . . . . . . . . . . . . . . , .... ' (" , "

~o ~,so t40 tao ~o ~oo 100 llO i~0 1~o t4o 15o
t. t

t,

E
~176

Fig. 2. Time evolution of the rms Mach


number M (--) and the pointwise local maxi-
mum of the Mach number (lul/c) a (---) for
L (a) case I, (b) case II, and (c) case III.

also relaxes to a value around 0.03. The amplitude of oscillations in both


compressive ratios for case III increases in time.

3.2. Energy Transfer


We consider here the effect of compressibility on energy-exchange
among the kinetic and internal components. The time variation of total
energy density E r is described by (2.3). The equation for kinetic energy
density EK is derived from (2.1) and (2.2). Subtracting (2.1) multiplied by
89 from (2.2) multiplied by us, we obtain

0 ~ (U~El,:) - u~ t3p 2

- (AD) + ( P D ) + (VD) (3.1)


Decaying Compressible Turbulence 9

1
C
l

~.g.

,so t~o ~o ~o ~6o 20o 100 110 t20 130 i.I0 150
[ l

:3

E
ctl

Fig. 3. Time evolution of the small-scale com-


30 3'2 3"4 3e 3'6 ,10
pression ratio rcs for (a) case I, (b) case II, and
t (c) case III.

The three terms on the right-hand side of (3.1) are called the advection
(AD), pressure-gradient (PD), and viscous-dissipation (VD) terms, respec-
tively. By taking the difference of (2.3) and (3.1), we obtain an equation for
the internal energy density El:

8 8 1
8t E~-- 8x~ (u~E~) - pA q- M2~ Pr Reo(y- 1) V2T

2
+ ~eo (S~Sij-' xA 2) (3.2)

Spatial averages of (2.3), (3.1), and (3.2) then give

d
dt <ET> = 0 (3.3)
10 Kida and Orszag

a
C

~2

c~.

=
~5o t~o t~0
t
1~o ~o

b
zoo
i10(] 110 12o
t.
130 140 /50

Fig. 4. Time evolution of the large-scale com-


pression ratio rcL for (a) case I, (b) case II,
30 3'2 3'4 3~6 3'a 40
t and (c) case III.

d 1
(EK> = <pA > - - ~ - - ( 4 < A 2 > + (1~o[2>)
dt tte o
= <PD> + <VD> (3.4)

and

d"-]l<fI>=
--<pzJ>-'k (4<~2> + <1~ol2>)

= -(PD> - <VD> (3.5)

Notice that both terms on the right-hand sides of (3.4) and (3.5) are
exactly the same but of opposite signs. This means that energy is exchanged
between the kinetic and internal components without changing the total
Decaying Compressible Turbulence 11

.,.,...,-"
.......

t50 t~o J
t
1oo llo L20
k
t00 140
I150

i] ...............
Ii
30
........
32
.....
31~
t
Q'6 3'8
I
40
Fig. 5. Time evolution of total (--), kinetic
(---), and internal (.-.) energies for (a) case I,
(b) case II, and (c) case III.

energy [see (3.3)]. The first term on the right-hand side of (3.4) describes
work done by pressure on a fluid element of unit volume. The internal
energy increases at the expense of kinetic energy where a fluid element con-
tracts (A <0). The energy-exchange is reversed where a fluid element
expands (A >0). The second term describes the viscous-dissipation of
kinetic energy, by which energy is always converted from kinetic to internal
components. For later convenience, we denote by e the negative of the
spatial average of (VD) and term it the energy-dissipation rate, i.e.,

~-= - ( V D ) =ec+~R
1
(4(A2) + (Io9[2)) (3.6)
= Reo
12 Kida and Orszag

o
~o

:7... /; ;--:....J,....,......-......?..,......-,......,
..... ~o
.~ r 84
c~

~o o \

i 150 ~;o ,~o ~;o ~o 200 ioo

t t

"d
mm~. ".. /''. ....................... ..../'",.,....../'"..,.. ........... .,.."

Fig. 6. Time evolution of ( P D ) (...) and


( V D ) (--) for (a) caseI, (b) caselI, and
?3o ~ ~'a
t (c) case III.

where

4
--- (L/2 ) (3.7a)
e c - 3 Re o

and

1
eR = oRe-- (10912) (3.7b)

are, respectively, the compressive and rotational components of the energy-


dissipation rate.
The time development of the total ( E r ) , kinetic ( E r ) , and internal
(El) energies for cases I, II, and III is plotted in Figs. 5a, 5b, and 5c,
Decaying Compressible Turbulence 13

71
c.

,=J

L
loo 1~0 120 130 140 150
t t

,u

Fig. 7. Decay of kinetic energy plotted using


linear-log scales for (a) case !, (b) case II, and
3O 32 3'4 ~'e 3'e 40
(c) case III. The straight lines show exponen-
t tial decay laws.

respectively. For cases I and II, the kinetic (internal) energy decreases
(increases) monotonically in time keeping the total energy constant. The
similar time variations are observed in the mean values of these two com-
ponents of energy for case III. But in this high-compressive ratio case they
oscillate regularly with period about 4.0, which results from collisions of
sound waves (KO). Anyway, in all the cases, the energy is transferred from
the kinetic to the internal components in the mean.
In Figs. 6a, 6b, and 6c we plot the time variations of ( P D ) (dotted
line) and ( V D ) (solid line) for cases I, II, and III, respectively. In cases I
and II, the VD term is dominant and changes smoothly in time. The PD
term is small in magnitude and oscillates around zero. Therefore the energy
decay is essentially by viscous dissipation. In case III, the roles of the VD
and PD interactions are reversed. The latter oscillates around zero with
amplitude much greater than the former. Because the PD interaction is
14 Kida and Orszag

c
T
150 ,;o 6o ~8o ~o 200 30 3'2 3~ ~6 ~'8 ,o
t t

Fig. 8. Characteristic time of exponential decay of energy. T= (din E/dt) -1. (a) Case I and
(b) case II.

dominant, the oscillation of ( P D ) synchronizes with that of (EK). The


phase of oscillation of ( P D ) is 90 ~ behind that of ( E K ) (cf. Fig. 5c).

3.3. Energy Decay Law


To see the decay law of kinetic energy in more detail, we plot the time
evolution of ( E r ) for cases I, II, and III in Figs. 7a, 7b, and 7c, respec-
tively. The data for cases I and II are fitted well by straight lines, implying
that the energy decays exponentially in time. During the simulations, the
energy decreases by factors of 10 and 4 for cases I and II, respectively. The
results for case III oscillate regularly while the mean value decreases a little
more slowly than exponentially in time.
In order to examine the decay law quantitatively we calculated the
reciprocal of the logarithmic derivative of the kinetic energy,

(dlnE~ -~ (3.8)
T=\ art /
In Figs. 8a and 8b we plot the values of T calculated numerically for cases I
and II, respectively. Although the fluctuations are not small, a systematic
dependence of T on time does not seem to exist. The horizontal lines show
the mean values of T, which is - 2 2 over 160 ~< t ~<200 for case I and - 7 . 0
over 32 ~< t ~<40 for case II. Constancy of T implies that the kinetic energy
decays exponentially in time as

Er(t) ocexp(-T) (3.9)


Decaying CompressibleTurbulence 15

The straight lines in Figs. 7a and 7b represent this exponential decay law.
The mean velocity squared (lu[ 2) shows a very similar behavior, namely,
it also decays exponentially in time (figures omitted).
It may be interesting to note that the characteristic decay time T is
comparable with a time scale that is constructed in terms of a large-scale
characteristic length L and rms velocity U. If we take the largest possible
length 2~ as L (recalling that the initial flow field is driven by a random
force with Fourier modes of Ik[ = 1) and the rms velocity at the initial time
(taken from Table I) as U, then we find L/U=21 for case I and 6 for
case II. q'hese values are very close to the decay time T. The constancy of
T implies that L changes in time in the same way as U, i.e., exponentially
in time.
The dissipation time scale td= (#/e)m ranges 2.0 and 5.5 over the
whole decaying period for case I, and 0.51 and 1.0 for case II. Thus, the
e 1-fold decay time T is several times longer than the dissipation time
scale td.
It should be mentioned here that exponential decay has also been
observed in moderate Reynolds number incompressible turbulence (Kida
and Murakami, 1987). Using their data, we confirm that T is more than
several times larger than td in the incompressible case also.

3.4. Compressive and Rotational Components

In order to study the effect of compressibility on turbulence it is useful


to decompose the kinetic energy into the compressive and rotational com-
ponents and examine the interaction among these two components and the
internal energy. This analysis has already been done for forced turbulence
(KO). Here, we apply the same method to decaying turbulence and
compare the results with the forced case.
In order to define the compressive and rotational components of
kinetic energy we introduce the new variable

w=x/pu (3.10)

and decompose it into three parts:

w=wR+wc+Wo (3.11)

where the rotational and compressive components, we and Wc, are, respec-
tively, divergence- and rotation-free, both of which vanish when spatially
averaged (see KO). The rotational, compressive, and mean components of
16 Kida and Orszag

kinetic energy are then defined by spatial averages of the squares of the
respective components of w as

<EKe> = 89 (3.12a)
<EKc>=89 (3.12b)
(E,<o> = ~<lwol
i 2> (3.12c)
The kinetic energy is the sum of these three components:

<EK> = (EKR> + ( E x c ) + (EKo) (3.13)


The equation describing the time variation of w can be derived from
(2.1) and (2.2) as

aW i aWi 1
at = -uJ-~xj-~ w'~
1 @ 2 Q
N ~ aXi + -Reo
- x//p axj (So.-- ~A6o) (3.14)

By taking a spatial average of the inner product of (3.14) with w~, ~ being
R, C, or O, we obtain
d
dt <EKe> = ( A D ) ~ + ( P D > ~ + (VD)~ (3.15)
where

,316a

(pD) =< 1 apw~i}


c~xl
(3.16b)

<VD >~ = < ( R e ; ~ p 0-0xj (Sij-lA6o))w~i> (3.16c)

The three terms on the right-hand side of (3.14) represent contributions to


the ~ component of kinetic energy from advection, pressure dilatation, and
the viscous-dissipation terms, respectively. It is easy to show that
(AD>R + <AD>c+ (AD>o = 0 (3.17a)
( P D > ~ + ( P D > c + ( P D > o = (PD> (3.17b)
<VD)R+ < V D ) c + < V D ) o = ( V D ) (3.17c)
Decaying Compressible Turbulence 17

Thus, the PD and VD interactions involve the compressive and rotational


components of kinetic energy as well as the internal energy, while the AD
interaction is closed in the former two components.
The time evolutions of these terms in the kinetic energy budget for
cases I, II, and III are plotted in Figs. 9a, 9b, 9c, respectively. Parts i, ii, iii
of Figs. 9a, 9b, and 9c are for the AD, PD, and VD terms, respectively.
Here the rotational, compressive, and mean components are drawn with
dotted, dashed, and solid lines, respectively. First of all, we note that all
the mean components, ( A D ) o , ( P D ) o , ( V D ) o , except for ( A D ) o
for case III (Fig. 9c-i), are very small in magnitude compared with the
rotational and compressive components.
For the AD term, the rotational and compressive components oscillate
periodically in time. The periods of the oscillations are almost constant but
the amplitudes decrease in time. The oscillations of the rotational and com-
pressive components are almost 180 ~ out of phase most of the time, so that
the energy is transferred back and forth between these two components.
Similar behavior was observed for forced turbulence (KO).
As for the PD term, the compressive component is dominant in all the
three cases. It oscillates around zero so that the direction of energy
exchange between the compressive and internal components alternates
periodically. The behavior of the PD term is also the same as for forced
trubulence (KO). The oscillations seen in ( A D ) and ( P D ) may be
related with interference between sound waves. Sound waves travel across
the periodic box roughly 7, 2, and 6 times during the simulations for
cases I, II, and III, respectively. Since there are two families of sound waves
which travel in opposite directions, the number of interactions (head-on
collisions) of sound waves may be estimated as 14, 4, and 12 times (double
the transit time) during the simulation periods of the respective cases.
These numbers of interactions are comparable with the numbers of oscilla-
tions in AD and PD terms, but a detailed comparison is necessary to
confirm the one-to-one correspondence between the oscillations and
interactions of sound waves.
The VD term is always negative and steadily converts kinetic energy
into internal energy. The ratio of the rotational and compressive com-
ponents of energy-dissipation rate is directly related to the small-scale com-
pressive ratio [see (2.15) and (3.6)]. Throughout the simulation, the rota-
tional component is dominant for cases I and II, while the compressive
component dominates for case III. As observed in the forced case (KO),
the series of negative peaks of ( V D ) synchronize with maxima of
kinetic energy (cf. Fig. 7c) so that (A 2) and (lucl 2) take large values
simultaneously. Thus the compressive components of large- and small-scale
motions are strongly correlated as in shock waves.

854/7/I-2
18 Kida and Orszag

r,- i!, i!:'. ~, J

7
i'J
(a-i)
150 ,~o ,~o ~o ~o ~oo
t

~T

(a-ii)

~5o ~o ~o ~o t~o 2oo


C
'o =
~o

s............,.... ....
...

/...."
./

~,~- ....-
7
//
.y-"
(a-iii)
.............

150 .I~0 Z70 I00 I00 2OO


I

Fig. 9. Time evolution of terms contributing to the kinetic energy budget for (a) case I,
(b) caselI, and (c) caselII. (i) (AD}~, (ii) (PD)~, (iii) (VD)~, where e = R , C, orO.
Solid, dotted, and dashed lines represent the mean, rotational, and compressive components,
respectively.
Decaying Compressible Turbulence 19

~- .

?- :

\x,/
(b-i)
3O 3'2 3'4 0'6 3'8
t
6"

3_
/'\ .,,
o!/: ...........................-.,,,,. ",,.,,,,;/"'.,.,,...//
:: ',,.... .......

(b-ii)
?
3O 3'2 0'4 3~ 3a 40
t

........ ..........................

/ .,....""~"
...-
.,-
..-"
~ . ....."'
///"

~?-
....... ,.'"'"
...,"
/
//"
(b-ill) /

Fig. 9. Continued.
v

DissipaLion Term *L0-' Pressure-DilaLaLion Term , 1 0 -~ AdveeLJon Term ' 1 0 -7


-~0 -~.0 -~ -~ o.o -e.o -Lo o.o Lo 2,0 ~.o 4.0 -25.P,-20.0-15.0-10.0-5.0 0.0 5.0 Lo.0 t5.0 20.0 25.0
! . ....i

.~2~\22.]]]iZ::;::::::: ........
~, ' . . . . . . at

- :'." . . . . . ,

-.-..'.
C]
0

!-
;o

L ~h

.=
Decaying Compressible Turbulence 21

4. SPECTRUM

4.1. Kinetic Energy Spectrum


The spectrum of kinetic energy is defined in terms of the Fourier
transform of w introduced in Sec. 3.4:

1 (,
'~(k) = ~ j w(x) exp[ - 2rcik 9x ] dx (4.1 )

where the integration is carried out over a periodic box. The wave number
k = (kl, k2, k3) has integer components. The square of the Fourier coef-
ficient

EK(k) -= 89I~(k)l e (4.2)

gives the kinetic energy spectral density, and the total kinetic energy is

( E K ) = ~ EK(k) (4.3)
k

where the summation is taken over -89 k2, k3<~89 The band-
averaged three-dimensional kinetic energy spectrum EK(k) is defined by

EK(k)= 89 ~ [~(k')l 2 (4.4)


k-- 1/2~< Ik't < k + 1/2

where k = 0, 1, 2,.., i2N and the total kinetic energy is


O/2)N
(EK)= ~.. EK(k) (4.5)
k=O

The kinetic energy spectra of the rotational and compressive components


are similarly defined in terms of the Fourier coefficients of the respective
components of w as

EKe(k)=89 ~ ]~(k')[ 2, a=R or C (4.6)


k-- 1/2~< Ik'[ < k + 1/2

The sum of these two spectra is the (total) kinetic energy spectrum:

EK(k) = E,,R(k) + EKe(k) (4.7)

Perspective views of isosurfaces of EK(k) for intermediate wave num-


bers at t = 40 for case II and at t = 150 for case III are shown in Figs. 10a
and 10b, respectively. [Similar shapes of isosurfaces of EK(k) are observed
22 Kida and Orszag

at other levels, too.] The surface levels are 10 -s for case II and 10 -11 for
case III. These levels are plotted by straight lines in Figs. l lb-ii and l lc-ii
for reference, noting the isotropic relation E~:(k)=4nk2Ex(k), Ikl=k.
Remember that the spectral density at the initial instant is isotropic in
wave number space for case II, while it is localized near the three wave-
number axes for case llI (see Fig. 8 in KO). As seen in Fig. 10a, the
isotropy in the energy spectral density is preserved 'throughout the simula-
tion for case II. On the other hand, the localization of the spectral density
near the three wave-number axes is not smeared out in case III at t = 150
after 5 Kolmogorov times [see (4.11) belowl have elapsed. This suggest
that the interactions among Fourier modes with wave numbers of different
directions are relatively weak in case III, where both the vorticity and the
rotational components of velocity are extremely small (see Table I).
In Figs. lla, lib, and llc, we plot three kinds of kinetic energy
spectra at several times for cases I, II, and III, respectively. (For the initial
spectra, see Fig. 9 in KO.) Solid, dotted, and dashed curves represent the
total, rotational, and compressive kinetic energies, respectively.
In case I, the compressive component is a few orders of magnitude
smaller than the rotational component over the whole wave-number range.
The rotational component and total energy spectra overlap in this graph.
At larger wave numbers both the components have exponential tails. Note
that the rms Mach number is quite low ( M = 0.27 at t = 160 and M = 0 . 1 1
at t = 200).
In case II, the rotational component is dominant at small wave num-
bers but the compressive component dominates at large wave numbers at

(b)
k3

f"

Fig. 10. Perspective reviews of kinetic energy spectral density (a) at t = 40 for case II and
(b) at t = 150 for case III. The isosurface levels are shown by straight lines in Figs. llb-ii and
llc-ii for the respective cases.
Decaying Compressible Turbulence 23

(a-i) (a-ii)
"o < '
R = 200 R = 200
'c
t = 160 t = 200
R~, : 29 R a = 13
_Xi = 0.27 M=0.11

k k

(b-t) (b-ii)
%
R=50 -- R:50
t:31 t:40
It~ : 34 E~ R~ = 17
M : 0.73 ~ M = 0.35
@o i
A
'= ......>,.
~-o

I0~ . . . . . . . . I'0'
k k

(c-i) (c-ii)
'c
R = I00 R = 100
'o
t : 120 t -= 150

E ~ .... R ~ = 40 R x -- 4.3
M = 0.07 M = 0.05
~. 9 ............ ..,...

~o.-"
~. .......

'o

'o
10 io'
k k

Fig. 11. Three-dimensional kinetic energy spectra for (a) case I, (b) case II, and (c) case III.
- - , E r ( k ) ; .--, ExR(k); ---, EKc(k ). The straight lines are the curves E K ( k ) = 4 ~ c k 2, where
c = 10 -8 for part b-ii and c = 10 -al for part c-ii.
24 Kida and Orszag

t = 31, at which M = 0.73. The abrupt change of the compressive compo-


nent of velocity across shock waves is responsible for raising the spectrum
at large wave numbers. As time increases, the compressive components
decay more rapidly at large wave numbers than the rotational component.s.
At small wave numbers, on the other hand, the decay of the compressive
component is less than the rotational component. At t = 40, at which
M=0.35, the rotational component dominates over the whole wave-
number range.
In case III, in contrast to cases I and II, the compressive component
is a few orders of magnitude larger than the rotational component over
the whole wave-number range. The compressive component cannot be dis-
criminated from the total energy spectrum.

4.2. Kolmogorov Similarity


It has been established for incompressible flow that energy spectra for
different kinds of turbulence collapse onto a single curve at large wave
numbers if the spectra are normalized according to Kolmogorov's
similarity law (Monin and Yaglom, 1975). We extended in KO the
Kolmogorov similarity analysis to compressible flows. There we introduced
a dissipation (Kolmogorov) wave number k(dK) by
k~K)= (p)1/2 (Reo3 e)1/4 (4.8)

and a characteristic energy spectrum Eo by

E(oK)_ 1 ( e ~ ~/4
( p ) 1/2 \ ~ e 05j (4.9)

The Kolmogorov wave numbers k f ) and k~m, and the characteristic


energy spectra Eo~c) and Eo~R) for the compressive and rotational com-
ponents are similarly defined in terms of ~c and ~n. Then, the normalized
energy spectrum F~, ~ = K, KR, or KC, is defined by

E~,(k ) = E(o~')F~,( + ) (4.10)

We found that the energy spectrum of the rotational component obeys the
Kolmogorov similarity law but the compressive component does not. The
kinetic energy spectrum, which is the sum of the rotational and com-
pressive components, therefore does not obey the similarity law. However,
if the rotational component is dominant, the kinetic energy spectrum
satisfies approximately the Kolmogorov similarity law. Here we report a
similar study for decaying flows.
Decaying Compressible Turbulence 25

Table II. Kolmogorov Wave N u m b e r s

Case kd k ~m

At the initial time


I 10 10 3.5
[I 9.9 9.2 7.1
III 4.6 -- 4.6

At the final time

I 6.0 6.0 2.3


II 6,9 6.9 2.9
III 2.6 -- 2.6

The Kolmogorov wave numbers at the initial and final times of the
simulations for the three cases are given in Table II. The smallness of ka
compared with the maximum wave number (=32) in the simulation
ensures that the error due to truncation of the Fourier components may be
negligible. The characteristic energy spectrum E0~) can be estimated using
the relation E~o~)=k~a~)/Re2, which follows from (4.8), (4.9), and ( p ) = 1.
The Kolmogorov time
1
tK= (Reo e) 1/2 (4.11)

represents the characteristic time of the smallest scale of motion. In


Table III we given the Kolmogorov time scale at the initial and final times
for the three cases. The Kolmogorov time tx increases in time as the
energy-dissipation rate decreases. Roughly speaking, the present simula-
tions extend for 13, 12, and 5 Kolmogorov times in cases I, II, and III,
respectively.
In Fig. 12 we plot the normalized three-dimensional kinetic energy
spectrum for caseI at (a) t = 1 6 0 (R~=29, M=0.29, rcs=0.007,

Table III. Kolmogorov Time

Case Initial Final

I 2.0 5.5
II 0.5 1.1
III 4.7 I5
26 Kida and Orszag

rCL=0.010) and (b) t = 2 0 0 ( R x = 13, M = 0 . 1 1 , r c s = 0 . 0 1 7 , rCL=0.029).


(See Fig. 10 in K O for the spectrum averaged over a forcing period before
t = 150, which is close to the initial energy spectrum.) The solid curve is the
normalized spectrum for incompressible flow (Kida and Murakami, 1987), 3
which was determined numerically as

Fx (~) ~ 8.4~ - 1.6 exp( - 4.9K) (4.12)

3 The same form of the normalized energy spectrum is observed for forced incompressible
turbulence (Kida et al., 1990).
"b
R = 200
t = 160

R x = 29
M = 0.27
c':
rCL -----0.016
rcs = 0.009

'oj

,2!
lO-L td
Normalized Wavenumber

- R = 200
-o, t = 200

"~ Rx = 13
~=c M = 0.II

~- rCL = 0.020
~,: :
9~% r c s = 0.007

(b) '~
'O l
i0"I i0~
Normalized Wavenumber

Fig. 12. Normalized three-dimcnsional kinctic energy spectrum at (a) t = 160 and (b) t = 200
for case I. The solid curve represents the normalized spectrum for a realization of incom-
pressible turbulence (Kida and Murakami, 1987).
Decaying Compressible Turbulence 27

R = 50
t=31
R~ = 34
M = 0.73
rCL : 0.050
rCS = 0.21

(a)
'c . . . . . . . . .
i0" 10~
Normalized ]Yavenumber

IR=50
-_c
t=34
R~ = 27
-\ jM=o58
~9 [rCL = 0.037
r c s 0 069

=
,%"

(b)

Normalized Wavenunaber

R=50
"c
t=40
R:, = 17
E-
M = 0.35
u :
rCL = 0.027
rn-'~
rcs -- 0.027

(c)
'o
\
IO" . . . . . . Ib'
Nor rn~liized t t a v e l u l m b ~ r

Fig. 13. Normalized three-dimensional kinetic energy spectrum at (a) t = 31, (b) t = 34, and
(c) t = 4 0 for case II. The solid curve is the same as that in Fig. 12.
28 Kida and Orszag

It is seen that the normalized spectrum agrees quite well with that for the
incompressible case over the whole wave-number range calculated. The
agreement is better at later times probably because the Mach number is
lower. There is no apparent correspondence between the smallness of the
compressive ratios and the closeness to the universal form of the spectrum
for these small values of compressive ratios (rcs ~<0.02, rcL < 0.03).
The normalized three-dimensional kinetic energy spectra at (a) t = 31
(R~ = 34, M = 0.73, rcs = 0.21, rcL = 0.050), (b) t = 34 (R~ = 27, M = 0.58,
rcs=0.069, rcL=0.037); and (c) t = 4 0 (R~=17, M=0.35, rcs=0.27,
rcL =0.027) for case II are plotted in Fig. 13. The spectrum deviates from
the universal form for incompressible flow (solid line) at large wavenum-
bers at early times. This is due to the strong excitation of the compressive
component for larger Mach numbers. However, the agreement is better at
later times as the Mach number decreases. The spectrum begins to deviate
from the universal form a t k/k d ~ 1 for M ~ 0.7 and at k/k d ~ 2 for M ~ 0.6,
but no substantial deviation is seen for M~0.35.
Now we examine similarity laws of the rotational and compressive
energy spectra separately. Relevant Reynolds numbers for the respective
components may be chosen as the rotational and compressive microscale
Reynolds numbers defined by
R(R) = ( 5 ~1/2
\ 3--~R] (P)(IURI2) (4.13a)
and
Rio,__(\3#ecJ
5 ''2 (P)([uc[2) (4.13b)

The values of these two Reynolds numbers at the initial and final times are
listed in Table IV. The rotational Reynolds number is dominant and nearly

Table IV. Compressive and Rotational Reynolds N u m b e r s

Case R~m R~c)

At the initial time


I 44 4.3
II 38 4.6
III -- 5.2

At the final time


I 13 2.5
II 16 2.5
III -- 3.2
Decaying Compressible Turbulence 29

equal to the Taylor microscale Reynolds number R~ (Table I) for cases I


and II, while the compressive Reynolds number dominates for case III. The
compressive Reynolds number for the former cases is small (R~c) =4-5).
The rotational Reynolds number for the latter case cannot be estimated
because the velocity and energy-dissipation rate of this component are too
small.
In Figs. 14a, 14b, and 14c we plot the normalized three-dimensional
energy spectra of the rotational component for case II at t = 31, 34, and 40,
respectively. The solid curves are again the universal form of the energy
spectrum for incompressible flow (Kida and Murakami, 1987). By
comparing them with Figs. 13, we find that the rotational kinetic energy
spectrum is closer to the universal form than the total kinetic energy
spectrum. The difference between the rotational and total energy spectra is
smaller for smaller M. As noted already (Fig. lla), these two spectra can
hardly be discriminated for case I ( M < 0.3).
In Figs. 15a, 15b, and 15c we plot the normalized compressive kinetic
energy spectra at several times for cases I, II, and III, respectively. The
spectral form is much different depending on the Mach number as well as
on the compressive ratios. We cannot find any similarity behavior of the
spectrum with respect to these parameters. However, we can conclude that
the spectrum of the compressive component of energy decays more slowly
at large wave numbers for larger Mach numbers.
The power spectrum of velocity and its rotational and compressive
components are defined in the same way as the kinetic energy spectrum. As
was done for the energy spectrum, we normalize these power spectra by the
characteristic power spectrum Eo/(p) and characteristic wave number kd.
The normalized power spectra of the total and compressive velocities are
obtained in the same way as the kinetic energy spectrum and agree almost
perfectly with the corresponding kinetic energy spectra for all the three
cases (detailed plots being omitted here). This is quite surprising because
the density fluctuation is not very small for case II. (The same observation
was made for forced turbulence in KO.)
There is, however, an interesting difference in the spectra of the rota-
tional components of w [see (3.9)] and the velocity u. In Fig. 16, we plot
the normalized rotational velocity spectrum at t = 31, 34, and 40 for case I.
As time increases and the Mach number decreases, the power spectrum of
velocity tends to approach the universal form for the incompressible tur-
bulence just as the spectrum of energy does. However, we can see that the
former is closer to the universal form than the latter (cf. Figs. 14a and 16a).
30 Kida and Orszag

t=31

~ ".

(a)
i0 -~ 10m
Normalized Y{avenurnber

t~34

z. :
'c.

'c
(b)

Normalized Wavenumber

t=40

~'c

(e)
10" lo~
N~3rmJhTed ~Yav~n~mb,.r

Fig. 14. N o r m a l i z e d t h r e e - d i m e n s i o n a l kinetic energy s p e c t r u m of the r o t a t i o n a l c o m p o n e n t


at (a) t = 31, (b) t = 34, a n d (c) t = 40 for case II. The solid curve is the same as t h a t in Fig. 12.
Decaying Compressible Turbulence 31

lo
o
o6

~~
@
o
oD
oo
o%
o\
o%
o%

(a)
o
i0 - ~
10 -1 1 10

a ~xo

%
%
\ o

(b) 10_io
.O-t 1 I0

o o
m

?o
o~

o%
o%
o %
o %
o %
o~
o
o
(c) %
tO -~ 1 10

Fig. 15. Normalized three-dimensional kinetic energy spectrum of the compressive compo-
nent. (a) [] at t = 170 (R~ = 23, M = 0.21, rcs = 0.008, rCL = 0.014) and O at t = 200 (R;. = 13,
M = 0 . 1 1 , rcs=0.017 , rcL=0.029) for caseI, (b) A at t = 3 1 (Ra=34, M = 0 . 7 3 , rcs=0.21,
rCL=0.050), [] at t = 3 4 (Rx=27, M = 0 . 5 8 , rcs=0.069, rCL=0.037) and O at t = 4 0
( R ; = 1 7 , M = 0 . 3 5 , rcs=0.27 , rcL=0.027 ) for caselI, and ( c ) [ ] at t = 1 2 0 (Rx=4.0,
M = 0 . 0 7 , rcs=0.997, rCL=0.83) and O at t = 1 5 0 (Rx=4.3, M = 0 . 0 5 , rcs=0.997,
rCL = 0.74) for case III.
32 Kida and Orszag

R:50
t:31

velocity

.~,o_

:[.!

(~)

- 1 ~'(v ' ' "" ' '" ' ' " ~ 'l"'v
Normalized Y/avenumber

I<= 50
t=34

i,o

(b)
-io" ' . . . . . . . f~ '
Normalized Wavenurnber

R=5o
t=40

&

(c)
-10-' ]O* '
Norrn~}ized wavenu tuber

Fig. 16. Normalized three-dimensional power spectra of the rotational component of


velocity at (a) t = 31, (b) t = 34, and (c) t = 40 for cases II. The solid curve represents the nor-
malized spectrum for a realization of incompressible turbulence (Kida and Murakami, 1987).
Decaying Compressible Turbulence 33

5. SUMMARY AND DISCUSSION


Here we studied the statistical properties of decaying compressible tur-
bulence by solving numerically the compressible Navier-Stokes equation
supplemented by the equation of state for an ideal gas. The initial flow field
is a fully developed turbulent state which was generated by a random
external force. The flow is confined in a periodic cube. The characteristics
of the flow depend on the microscale Reynolds number R;~, the rms Mach
number M, and the small- and large-scale compressive ratios rcs and rcc.
We obtained the following results.
(1) The kinetic energy (and also the velocity squared) decays
exponentially in time. This decay law is clearly seen for the case that the
compressive ratios are so small that the energy exchange between the
kinetic and internal components through the pressure-dilatation interaction
may be negligible compared with the energy dissipation by viscosity. If the
Compressive ratios are large, the kinetic energy exhibits a decaying oscilla-
tion resulting from a periodic energy exchange between the compressive
component of kinetic energy and the internal energy (Passot and Pouquet,
1987). The exponential decay law of energy has also been observed
for moderate Reynolds number incompressible turbulence (Kida and
Murakami, 1987).
(2) The kinetic energy spectrum (as well as the power spectrum of
velocity) of the rotational component obeys the Kolmogorov similarity law
at least for Mach numbers M up to about 0.6. On the other hand, the
spectrum of the compressive component varies significantly at larger wave
numbers depending on the Mach number and the compressive ratios. This
suggests that even if the Mach number is moderate (M~0.5), the
compressive and rotational components behave rather independently of
each other as far as the modal distribution of energy is concerned.
(3) In the free decay of turbulence, the microscale Reynolds number
and the rms Mach number decrease in time but the two compressive ratios
seem to approach universal values. Because of the shortness of the simula-
tion period, however, it is not clear from the present numerical results
whether an asymptotic state, if it exists, is unique or not (Sarkar et al.,
1991; Erlebacher et aL, 1990).
Our results concern statistical properties of compressible turbulence
with relatively small Reynolds number (R~ <40) and rms Mach number
(M~0.8). These parameter values seem to be the upper limits reachable
with reasonable accuracy using simulations with 643 collocation points. We
have shown that the Kolmogorov similarity law does still hold for the rota-
tional component of the flow. However, higher-resolution simulations are

854/7/1-3
34 Kida and Orszag

necessary to realize the Kolmogorov inertial range spectrum and to


examine interactions among Fourier components for higher Reynolds and
Mach number turbulence.

ACKNOWLEDGMENTS
This work was supported by the Office of Naval Research under
contract No. N00014-82-C-0451, the Air Force under grant No. AFOSR-
90-0124 and contract No. F33657-91-C-2147, and DARPA under con-
tract No. N00014-86-K-0759. The computations were performed on the
Cray YMP at the Pittsburgh Supercomputer Center, which is supported by
the National Science Foundation.

REFERENCES
Erlebaeher, G., Hussaini, M. Y., Kreiss, H. O., and Sarkar, S. (1990). The analysis and
simulation of compressible turbulence, Theor. Comp. Fluid Dyn. 2, 73.
Kida, S., and Murakami, Y. (1987). Kolmogorov similarity in freely decaying turbulence,
Phys. Fluids. 30, 2030-2039.
Kida, S., and Orszag, S. A. (1990a). Energy and spectral dynamics in forced compressible
turbulence, J. Sci. Comput. 5, 85-125.
Kida, S., and Orszag, S. A. (1990b). Enstrophy budget in a decaying compressible turbulence,
J. Sci. Cornput. 5, 1-34.
Kida, S., Murakami, Y., Ohkitani, K., and Yamada, M. (1990). J. Phys. Soc. Jpn. 59, 4323.
Passot, T., and Pouquet, A. (1987). Numerical simulation of compressible homogeneous flows
in the turbulent regime, J. Fluid Mech. 181, 441-466.
Sarkar, S., Erlebacher, G., Hussaini, M. Y., and Kreiss, H. O. (1991). The analysis and
modeling of dilatational terms in compressible turbulence, J. Fluid Mech. 227, 473-493.
Zeman, O. (1990). Dilatation dissipation: The concept and application in modeling com-
pressible mixing layers, Phys. Fluids A2, 178-188.

You might also like