Sunteți pe pagina 1din 17

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3, pp.

147163 (2007)

TURBULENT FLOW SIMULATIONS USING THE MACCORMACK AND


THE JAMESON AND MAVRIPLIS ALGORITHMS COUPLED WITH THE
CEBECI AND SMITH AND THE BALDWIN AND LOMAX MODELS IN
THREE-DIMENSIONS
Edisson Svio de Ges Maciel
CNPq Researcher, Rua Demcrito Cavalcanti, 152, Afogados, Recife, Pernambuco, Brazil, 50750-080
E-Mail: edissonsavio@yahoo.com.br
ABSTRACT: In the present work, the MacCormack scheme and the Jameson and Mavriplis scheme are implemented
with the Cebeci and Smith model and the Baldwin and Lomax model to perform turbulent flow simulations in a threedimensional space. The Navier-Stokes equations in conservative and integral forms are solved, employing a finite
volume formulation and a structured spatial discretization. The MacCormack scheme is a predictor/corrector method
which performs coupled time and space discretizations, while the Jameson and Mavriplis algorithm is a symmetrical
scheme and its time discretization is performed by a Runge-Kutta method. Both schemes are second order accurate in
space and time and require artificial dissipation to guarantee stability. The steady state problem of the supersonic
turbulent flow along a ramp is studied. The results have demonstrated that both models predict satisfactorily the
boundary layer separation region formed at the compression corner, reducing, however, its extension in relation to the
laminar solution, as expected.
Keywords: Navier-Stokes equations, MacCormack algorithm, Jameson and Mavriplis algorithm, Cebeci and Smith
turbulence model, Baldwin and Lomax turbulence model.

dimensional space. The scheme was initially


developed using a finite difference technique. The
method involved two steps: a predictor step and a
corrector step. In the predictor step, the derivatives
of the flux terms were calculated with forward
spatial discretization operators and in the corrector
step, these derivatives were calculated with
backward spatial discretization operators.
Jameson and Mavriplis (1986) emphasized the
substantial cost reduction in the calculations of the
Euler equation solutions. The method proposed by
Jameson, Schmidt and Turkel (1981) had been
proved to possess robustness, good accuracy and
sufficient sophistication for more complete
applications. The objective was to apply the scheme
to geometries like wing-fuselage, involving
engines, missiles and other typical components, to
represent a whole airplane. The work emphasized
the use of triangular cells which allowed more
flexibility in the description of complex geometries
and made the mesh generation process less
expensive. The fluid movement equations were
spatially discretized in an unstructured context. The
scheme used a finite volume formulation with
properties determined at the cell centroids.

1. INTRODUCTION
The development of aeronautical and aerospace
projects requires hours of wind tunnel testing. It is
necessary to minimize such wind tunnel type of test
because of the growing cost of such tests. In Brazil,
there is a lack of wind tunnels of great capacity for
generating supersonic flows or even high subsonic
flows. Therefore, Computational Fluid Dynamics,
(CFD), techniques are receiving much attention in
the aeronautical industry. Analogous to wind tunnel
modelling, the numerical methods determine
physical properties in discrete points of the spatial
domain. Hence, the aerodynamic coefficients of lift,
drag and momentum can be calculated.
Initially, non-upwind schemes were developed to
simulate flow over simple and complex geometries
owing to their simplicity in numerical
implementation.
Predictor-corrector
and
symmetrical schemes were the most employed
algorithms from the 60s to 80s. Some of them are
reported below.
MacCormack (1969) developed a numerical method
which is second order accurate in space and time to
solve the Navier-Stokes equations in a twoReceived: 19 Feb. 2007; Accepted: 1 Apr. 2007
147

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

Artificial dissipation operators were constructed to


guarantee second order spatial accuracy of the
scheme, except in the proximities of shock waves in
which the accuracy was reduced to the first order
(Jameson, Schmidt and Turkel, 1981). The time
integration used a Runge-Kutta method of five
stages.
Maciel (2005a) performed a study involving the
MacCormack (1969) and the Jameson and
Mavriplis (1986) schemes in a three-dimensional
space. The Euler and the Navier-Stokes equations
were solved and the physical problems of the
supersonic flow along a ramp and the cold gas
hypersonic flow along a diffuser were studied. The
results have demonstrated that both algorithms have
described appropriately the flow field. The shock at
the ramp is well detected and the shock interference
in the diffuser problem is appropriately solved by
both schemes. The Cp distributions generated by
the algorithms in the ramp and in the diffuser
problems describe reasonably the shock and the
expansion fan in the inviscid case, but losing
quality in the viscous case.
There is a practical necessity in the aeronautical
industry and other fields for a model capable of
calculating separated turbulent compressible flows.
With the available numerical methods, researchers
seem to be able to analyze several separated flows,
three-dimensional in general, when an appropriate
turbulence model is employed. Simple methods
such as the algebraic turbulence models of Cebeci
and Smith (1970) and of Baldwin and Lomax (1978)
supply satisfactory results with low computational
cost and allow detection of the main features of the
turbulent flow.
Maciel (2006a) performed a comparison between
the Cebeci and Smith (1970) and the Baldwin and
Lomax (1978) models in relation to solution quality
and numerical accuracy, in a two-dimensional
space. The numerical algorithms of MacCormack
(1969) and of Jameson and Mavriplis (1986) were
implemented, using finite volumes and structured
spatial discretization, to perform the numerical
experiments. The Reynolds average Navier-Stokes
equations were solved. The steady state supersonic
flow along a ramp was studied. The results have
demonstrated that the Cebeci and Smith (1970)
model yielded better quality characteristics and
more critical solutions than the Baldwin and Lomax
(1978) model when the MacCormack (1969)

scheme was used. But when the Jameson and


Mavriplis (1986) scheme was studied, no
meaningful differences were perceptible.
In the present work, the MacCormack (1969) and
the Jameson and Mavriplis (1986) schemes are
implemented, using finite volumes and structured
spatial discretization, to solve the Reynolds average
Navier-Stokes equations in the three-dimensional
space for application to the problem of the
supersonic flow along a ramp. The implemented
schemes are second order accurate in both space
and time. It is necessary to introduce a dissipation
operator to guarantee the numerical stability of the
schemes and therefore, the Mavriplis (1990) and the
Azevedo (1992) models are implemented. The
algorithms are accelerated to the steady state
solution using a spatially variable time step. The
results have demonstrated that the Jameson and
Mavriplis (1986) scheme yields more severe Cp
distributions although the biggest value of Cp is
obtained with the MacCormack (1969) scheme
using the Cebeci and Smith (1970) model. The
shock angle of the oblique shock wave is best
predicted by the MacCormack (1969) scheme using
the Cebeci and Smith (1970) model. This work is a
continuation of the study initiated by Maciel
(2005a).
2. NAVIER-STOKES EQUATIONS
The
Navier-Stokes
equations
in
integral
conservative form, employing a finite volume
formulation and a structured spatial discretization
for three-dimensional simulations, can be written
as:

Q / t + 1 / V PdV = 0
v

(1)

where V is the cell volume, which corresponds to an


hexahedron in the three-dimensional space; Q is
the vector of conserved variables; and
r
P = (E e E v )i + (Fe Fv ) j + (G e G v )k
represents the complete flux vector in Cartesian
coordinates, with the subscript e related to the
inviscid contributions or the Euler contributions and
v to the viscous contributions. These components
of the complete flux, as well as the vector of
conserved variables, are described below:

148

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

v
w
u
u

uw
u 2 + p
uv

Q = v , Ee = uv , Fe = v 2 + p , Ge = vw ,
w
vw
w2 + p
uw

(e + p )u
(e + p)v
(e + p) w
e

(2)

yx
xx

1
1

xy
yy
Ev =
, Fv =
and
Re
Re

xz
yz

xxu +xyv +xzwqx


yxu +yyv +yzwqy

(3)

zx

1
zy
Gv =

Re

zz

zxu +zyv +zzwqz

In these equations, the components of the viscous stress tensor are defined as:

xx = 2(M + T ) u x 2 3(M + T )(u x + v y + w z) , xy = (M + T )(u y + v x) ;

(4)

xz = (M + T )(u z + w x) , yy = 2(M + T ) v y 2 3(M + T )(u x + v y + w z) ;

(5)

zz = 2(M + T ) w z 2 3(M + T )(u x + v y + w z) and yz = (M + T )(v z + w y) .

(6)

The components of the conductive heat flux vector are defined as:

q x = ( M Pr d + T Pr d T ) ei x , q y = ( M Pr d + T Pr d T ) ei y ;

(7)

q z = ( M Pr d + T Pr d T ) ei z .

(8)

The quantities that appear above are described as


follows: is the fluid density, u, v and w are the
Cartesian components of the flow velocity vector in
the x, y and z directions, respectively; e is the total
energy per unity volume of the fluid; p is the fluid
static pressure; ei is the fluid internal energy; the
values represent the components of the viscous
stress tensor; Pr d is the laminar Prandtl number,
which assumes a value of 0.72 in the present
simulations; Pr dT is the turbulent Prandtl number,
which assumes a value of 0.9; the q values represent
the components of the conductive heat flux; M is
the fluid molecular viscosity; T is the fluid
turbulent viscosity; is the ratio of specific heats at
constant pressure and volume, respectively, which
assumes a value 1.4 for the atmospheric air; and Re
is the Reynolds number of the viscous simulation,
defined by:

Re = u

REF

(9)

where uREF is a characteristic flow velocity and


l is a configuration characteristic length. The
molecular viscosity is estimated by the empiric
Sutherland formula:
M = bT 1 2 (1 + S T )

(10)

where T is the absolute temperature (K),


b=1.458x10-6 kg/(m.s.K1/2) and S=110.4 K, of the
atmospheric air in the standard atmospheric
conditions (Fox and McDonald, 1988).
The
Navier-Stokes
equations
were
nondimensionalized in relation to the freestream
density, , the freestream speed of sound, a ,
and the freestream molecular viscosity, , for the
studied problem. To allow the solution of the matrix
system of five equations with five unknowns
149

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

described by equation (1), the state equation of


perfect gases is employed, in its two versions, as
presented below:
p = ( 1) e 0.5 (u 2 + v 2 )

or

p = RT

(11)

i, j , k

(P S )

+ PS

i, j, k

r r
(P S )

( Pr Sr )
+
i, j, k
i , j 1 / 2, k

i, j, k

( )

i , j + 1 / 2, k

r r
+ (P S )

( )

+ PS

i 1 / 2, j , k

(14)
i, j , k 1 / 2
i, j, k + 1 / 2

r
where, for example, S i , j 1 / 2 , k has the same
r
direction as ni , j 1 / 2 , k and magnitude equals to the
S i , j 1 / 2 ,k area. The half indexes indicate flux
calculated at the respective faces or cell surfaces.
By discretizing space and time, according to the
Lax-Wendroff method, dividing the resultant
algorithm in two time integration steps (predictor
and corrector) and adopting a forward spatial
discretization in the predictor step and a backward
spatial discretization in the corrector step, it is
possible to obtain the MacCormack (1969)
algorithm, on a finite volume context, as follows:

i, j , k

with the n vector pointing outward of the


respective flux area. Details of the definition of a
computational cell and of the calculation of its
volume, the flux areas and the normal to the flux
faces are found in the work of Maciel (2002,
2006b).
The time marching using the Euler explicit method,
when applied to equation (12), leads to the
following expression:
z

i + 1 / 2, j , k

Applying the Green theorem to equation (1) and


adopting a structured notation to the fluid and flow
quantities, it is possible to write:
r r
(12)
Q
t = 1 V
(P n)
dS
S
i, j , k

Qn + 1 = Qn
i, j, k

3. EXPLICIT NUMERICAL SCHEME OF


MACCORMACK (1969)

i, j , k

i , j ,k

In the surface integral discretization, this equation


can be rewritten as:

with R being the specific gas constant. For


atmospheric air, the value is 287 J/(kgK).

i, j , k

r
S ( P n ) i , j ,k dS i , j ,k (13)

Qin, +j ,1k = Qin, j ,k t i , j ,k Vi , j ,k

Predictor Step:

(E E )n
V
S
+
Qn = t
+ (F F )n
+ (G G )n

i, j, k
i, j, k i, j, k e v xi, j 1/ 2, k
e v yi, j 1/ 2, k
e v zi, j 1/ 2, k
i, j 1/ 2, k

i, j, k

S
+
+ (F F )n
+ (G G )n

(Ee Ev )nx
e
v
z
e
v
y
1
/
2
,
,
i
j
k
1
/
2
,
,
+
i+
jk
i +1/ 2, j, k
i +1, j, k i +1/ 2, j, k

S
+
+ (F F )n
+ (G G )n

(Ee Ev )nx
i, j +1/ 2, k
e
v
z
e
v
y
i, j +1/ 2, k
i, j +1/ 2, k
i, j +1/ 2, k

i, j +1, k

S
+
+ (F F )n
+ (G G )n
(Ee Ev )nx
e v yi 1/ 2, j, k
e v zi 1/ 2, j, k
i 1/ 2, j, k
i 1/ 2, j, k

i, j, k

S
+
+ (F F )n
+ (G G )n

(Ee Ev )nx
i, j, k 1/ 2
e
v
z
e
v
y
,
,
1
/
2
i
j
k

,
,
1
/
2
i
j
k

i, j, k 1/ 2

i, j, k

S
+ (F F )n
+ (G G )n

(Ee Ev )nx
,
,
,
1
/
2
i
j
k
e
v
z
e
v
y
+
i, j, k +1/ 2
i, j, k +1/ 2
i, j, k +1/ 2

i, j, k +1
.
Q n +1 = Qn + Qn
i, j, k
pi, j, k i, j, k
150

(15)

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

Corrector Step:

(E E )n
Q n +1 = t
+ (F F )n
+ (G G )n
+
S
V

i, j 1/ 2, k
ci, j, k
i, j, k i, j, k e v xi, j 1/ 2, k
e v yi, j 1/ 2, k
e v zi, j 1/ 2, k

i, j 1, k

S
+ (F F )n
+ (G G )n
+
(Ee Ev )nx

i +1/ 2, j, k
e v yi +1/ 2, j, k
e v zi +1/ 2, j, k
i +1/ 2, j, k

i, j, k
p

E
E
n
(
)
(
)
(
)
S
+
+ F F n
+ G G n
e v x
i, j +1/ 2, k
e v yi, j +1/ 2, k
e v zi, j +1/ 2, k
i, j +1/ 2, k

i, j, k
p

+
(

)
+
(

)
+
(

)
E
E
n
S
F
F
n
G
G
n
e v x
i 1/ 2, j, k
e v yi 1/ 2, j, k
e v zi 1/ 2, j, k
i 1/ 2, j, k

i 1, j, k
p

+ (F F )n
+ (G G )n
+
S
(Ee Ev )nx

i, j, k 1/ 2
e v yi, j, k 1/ 2
e v zi, j, k 1/ 2
i, j, k 1/ 2

i, j, k 1
p

+ (F F )n
+ (G G )n
S
,
(Ee Ev )nx

i, j, k +1/ 2
e v yi, j, k +1/ 2
e v zi, j, k +1/ 2
i, j, k +1/ 2

i, j, k

n+1

n+1

Qin, +j,1k = 0.5 Qin, j,k + Qpi, j,k + Qci, j,k

(16)

As in the two-dimensional studies (Maciel, 2006c),


an explicit addition of an artificial dissipation
operator is needed in the approximation of a second
and a fourth derivative (Maciel, 2006d) to guarantee
numerical stability of the scheme at the proximities
of nonlinearities of the flow, such as shock waves,
and due to even-odd uncoupled solutions, which
generate independent solutions in even and odd
cells, that are physically incorrect. This operator is
subtracted from the flux terms in the RHS in the
corrector step. The artificial dissipation models
employed in conjunction with the MacCormack
(1969) scheme are based on the the work of
Mavriplis (1990) and Azevedo (1992), and details of
these implementations are found in Maciel (2002,
2005b, 2005c and 2006b).
In the calculation of the viscous flux vectors, the
derivatives present in Eqs. (4) to (8) are considered
constants in a given cell and are calculated in terms
of the surface integral of the variable along the
faces of the control volume (Green theorem).
Details of this implementation are found in Maciel
(2002, 2005b, 2005c and 2006b). The value of the
u x derivative at the (i,j-1/2,k) interface, for

cell (i,j,k) in the predictor step and as the value of


u x of the cell (i,j-1,k) in the corrector step.
4. EXPLICIT NUMERICAL SCHEME OF
JAMESON AND MAVRIPLIS (1986)
To obtain the Jameson and Mavriplis (1986)
scheme on a structured and finite volume context,
Eq. (12) should be rewritten as:
d (V

i, j, k

r r
(P S )

i, j, k

r r
r r
) dt + ( P S )
+ (P S )
+

i , j 1 / 2, k
i + 1 / 2, j , k

i , j + 1 / 2, k

r r
(P S )

r r
+ (P S )

i, j, k 1 / 2

i 1 / 2, j , k

r r
+ (P S )

i, j, k + 1 / 2

= 0

(17)

where:
r r
(P S )

i , j 1 / 2, k

=E E
v
e

i , j 1 / 2, k

i , j 1 / 2, k

+ (Fe Fv )i , j 1 / 2, k S yi , j 1/ 2 ,k

example, is adopted as the value of u x of the

+ G G

151

v i , j 1 / 2, k

i , j 1 / 2, k

(18)

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

represents the flux at the (i,j-1/2,k) interface, for


example. The inviscid and viscous flux vectors in
each flux interface are implemented using the
arithmetical average of the primitive variables in
each face; in other words, for the flux interface (i,j1/2,k), the primitive variables are determined by
calculating the arithmetical average of their values
at the cells (i,j-1,k) and (i,j,k).
As for the calculation of the viscous flux vectors,
the derivatives present in Eqs. (4) to (8) are
considered constants in a given cell and are
calculated in terms of the surface integral of the
variable along the faces of the control volume
(Green theorem), as mentioned in the MacCormack
(1969) scheme. The value of the u x derivative
at the (i,j-1/2,k) interface, for example, in the
Jameson and Mavriplis (1986) scheme, is
determined by the arithmetical average of the
values of u x related to the cells (i,j-1,k) and
(i,j,k).
The spatial discretization proposed by the authors is
equivalent to a symmetrical scheme with second
order accuracy, on a finite difference context. The
introduction of an artificial dissipation operator D
is necessary to guarantee the numerical stability of
the scheme in the presence of, for example, oddeven
uncoupled
solutions
and
nonlinear
instabilities, like shock waves. In both schemes of
MacCormack (1969) and Jameson and Mavriplis
(1986), two artificial dissipation models were
implemented: the first one based on the work of
Mavriplis (1990) and the second one based on that
of
Azevedo (1992).
Details
of
these
implementations are found in Maciel (2002, 2005b,
2005c and 2006b). Equation (17) is rewritten as:
dt + C (Q
d V
Q
) D (Q
) = 0

i, j , k
i, j, k
i, j, k i, j, k

second order accuracy, and can be represented in


general form as:
( 0)
( n)
=Q
Q
i, j , k
i, j , k
(k 1)
(k )
( 0)
( m)
=Q
t
D Q
Q
V
C Q

i, j , k
i, j , k
k i, j , k
i, j , k i, j, k
i, j , k
(n + 1)
(k )
=Q
Q
i, j , k
i, j , k

(21)
where k = 1,...,5; m = 0 until 4; 1 = 1/4, 2 = 1/6,
3 = 3/8, 4 = 1/2 and 5 = 1. Swanson and
Radespiel (1991) suggested that the artificial
dissipation operator should only be evaluated in odd
stages when the Navier-Stokes equations were
solved (m = 0, m = 2 and m = 4). This procedure
aims at CPU time economy and also better
smoothing of the numerical instabilities of the
discretization based on the hyperbolic/parabolic
characteristics of the Navier-Stokes equations.

5. TURBULENCE MODELS
5.1 Turbulence model of Cebeci and Smith
(1970)
The problem of the turbulent simulation is in the
calculation of the Reynolds stress. Expressions
involving velocity fluctuations, originating from the
average process, represent six new unknowns.
However, the number of equations remains the
same and the system is not closed. The modeling of
turbulence is to develop approximations to these
correlations. When calculating the turbulent
viscosity according to the Cebeci and Smith (1970)
model, the boundary layer is divided into internal
and external layers.
Initially, the kinematic viscosity (w) at wall and
the shear stress (xy,w) at wall are calculated. After
that, the boundary layer thickness (), the linear
momentum thickness (LM) and the boundary layer
tangential velocity (VtBL) are calculated. Then, the
normal distance from the wall (N) to the studied cell
is calculated. The N+ term is obtained from:

(19)

with:
r r
r r
+ (P S )
+
) = ( P S )

i, j, k
i , j 1 / 2, k
i + 1 / 2, j , k
r r
r r
(P S )
+ (P S )
+

C (Q

i , j + 1 / 2, k

r r
(P S )

i, j, k 1 / 2

i 1 / 2, j , k

r r
+ (P S )

+
i, j, k + 1 / 2

N + = Re

(20)

xy , w

N
w

(22)

where w is the wall density. The van Driest


damping factor is calculated by:

being the flux integral of the cell (i,j,k).


The time integration is performed using a hybrid
explicit Runge-Kutta method of five stages, with

D = 1 e
152

( N

/ w w / / A+

(23)

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

= 0.0168 , A0+ = 26 , C cp = 1.6 , C Kleb = 0.3

with A + = 26 and w is the molecular wall


viscosity. After that, the normal gradient of the
tangential velocity to the wall ( dVt dN ) is
calculated and the internal turbulent viscosity is
given by:

= Re (ND) 2 dVt dN

and C wk = 1 . FKleb is the intermittent function of


Klebanoff given by:

FKleb ( N ) = 1 + 5.5(C Kleb N N max )

maximum velocity value in the boundary layer case.


For free shear layers,

where is the von Krman constant, which has a


value of 0.4. The intermittent function of Klebanoff
is calculated for the external viscosity by:
g

Kleb

( N ) = 1 + 5.5(N )6

U dif =

(25)

Te

BL LM

Kleb

Ti

(26)

(27)

Te

7.1

lmix = N 1 e N

7.2

and

mix

A0+

In the external layer,


(29)

with:
= MIN N F ; C N U 2 / F and
max max wk max dif max

= 1 MAX l
N mix

max

+ v2

N = N max

(32)

Initial condition

Boundary conditions

8. RESULTS

wake

max

There are basically three types of boundary


conditions: solid wall, entrance and exit. These
conditions are implemented in special cells named
ghost cells and details of these implementations are
available in the work of Maciel (2002, 2005a,
2005b, 2005c and 2006b).

(28)

Te = Ccp Fwake FKleb ( N , N max / CKleb )

) ( u

Values of freestream flow are adopted as initial


condition, in the whole calculation domain of the
physical ramp problem (Jameson and Mavriplis,
1986; Maciel, 2002, 2005b, 2005c, 2006b, 2006c
and 2006d).

To the calculation of the turbulent viscosity


according to the Baldwin and Lomax (1978) model,
the boundary layer is again divided into internal and
external layers.
In the internal layer,
Ti

+ v2

7. INITIAL AND BOUNDARY CONDITIONS

5.2 Turbulence model of Baldwin and Lomax


(1978)

= l 2

The basic idea of this procedure consists in keeping


constant the CFL number in all computational
domain, allowing, hence, the use of appropriate
time steps in each specific mesh region during the
convergence process. Details of the present
implementation are found in the work of Maciel
(2002 and 2006b).

Finally, the turbulent viscosity is chosen from the


internal and the external viscosities:

= MIN ( , )

(u

6. SPATIALLY VARIABLE TIME STEP

With it, the external turbulent viscosity is calculated


by:

= Re(0.0168) Vt

(31)

is the vorticity vector magnitude and U dif is the

(24)

Ti

6 1

Tests were performed by using a microcomputer


with the processor AMD ATHLON XP 2600+
running at 1.91 GHz and 512 Mbytes of RAM
memory. As the interest of this work is the steady
state solutions, one needs to define a criterion to
check convergence to the steady state. The criterion
adopted in this work was a reduction of four orders
of magnitude of the maximum residue in the
domain, a typical criterion in the CFD community.

(30)

N max is the value of N where l mix reached its


maximum value and lmix is the Prandtl mixture
length. The constant values are: = 0.4 ,
153

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

The residue to each cell was defined as the


numerical value obtained from the discretized
conservation equations. As there are five
conservation equations to each cell, the maximum
value obtained from these equations is defined as
the residue of this cell. Thus, this residue is
compared with the residue of the others cells,
calculated as the same way, to define the maximum
residue in the domain. The downstream and
longitudinal plane angles were set equal to 0.0. In
the numerical experiments performed in this work,
the artificial dissipation model of Azevedo (1992)
was used.
The ramp problem is a supersonic flow hitting a
ramp with 20 of inclination. It generates a shock
wave and an expansion fan. The ramp configuration
in the xy plane is shown in Fig. 1. Its spanwise
length is 0.25 m. The mesh used in the simulations
has 37,260 hexahedra and 42,700 nodes in a
structured discretization of the calculation domain.
This mesh is the equivalent of, in finite differences,
one that is composed of 61 points in the direction,
70 points in the direction and 10 points in the
direction. An exponential stretching of 10% in the
direction was employed.

8.1

Fig. 2

Density contours (M-L).

Fig. 1

Fig. 3

Density contours (JM-L).

Laminar results

Fig. 2 and Fig. 3 exhibit the density field obtained


by the schemes of MacCormack (1969) and
Jameson and Mavriplis (1986) for laminar flow.
The Jameson and Mavriplis (1986) solution gives
denser field than the MacCormack (1969) solution.

Ramp configuration in the xy plane.

The freestream Mach number adopted as initial


condition to this simulation was 2.0, characterizing
a supersonic flow. The Reynolds number was
estimated to be 161,257 at a flight altitude of
20,000 m and l = 0.0437 m, based on the work of
Fox and McDonald (1988).

Fig. 4 and Fig. 5 show the pressure contours


obtained by the schemes of MacCormack (1969)
and Jameson and Mavriplis (1986) respectively. The
pressure field generated by the scheme of Jameson
and Mavriplis (1986) is more severe and critical
than the one generated by that of MacCormack
(1969). In other words, the scheme of Jameson and
Mavriplis (1986) is more conservative in that it
predicts a more severe pressure field. The pressure
field generated by the scheme of MacCormack
154

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

(1969) predicts a weaker shock before the ramp, at


approximately x = 0.08 m. This shock is due to the
increase in the thickness of the boundary layer,
originated from flow separation, resulting in a
circulation bubble as highlighted in Fig. 8.

Fig. 4

Fig. 5

streamlines highlight this separated flow close to


the wall, as well as the formation of a circulation
bubble. The separation region detected by the
scheme of MacCormack (1969) is longer than the
corresponding one detected by that of Jameson and
Mavriplis (1986). Moreover, due to its longer
region, a thicker boundary layer is characterized.
This effect tends to yield a weaker shock before the
ramp, as mentioned in the pressure field analyses
using the MacCormack (1969). The scheme of
Jameson and Mavriplis (1986), although also
detects boundary layer separation close to the ramp,
generates a smoother detachment when compared to
that observed in the scheme of MacCormack
(1969), without capturing the weaker shock ahead
of the ramp.

Pressure contours (M-L).

Fig. 6

Mach contours (M-L).

Fig. 7

Mach contours (JM-L).

Pressure contours (JM-L).

Fig. 6 and Fig. 7 exhibit the Mach number contours


obtained by both schemes. The Mach number is
more intense and has a better resolution, in the
solution obtained by the scheme of Jameson and
Mavriplis (1986).
Fig. 8 and Fig. 9 show the velocity vector fields
close to the ramp wall and the respective
streamlines obtained by the schemes of
MacCormack (1969) and Jameson and Mavriplis
(1986). Regions of separated flow are obtained in
the solutions generated by both schemes. The
155

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

Fig. 8

Velocity field and streamlines (M-L).

Fig. 10 Density contours (M-CS).

Fig. 9

Velocity field and streamlines (JM-L).

Fig. 11 Density contours (JM-CS).

Fig. 12 and Fig. 13 exhibit the pressure fields


obtained by the schemes of MacCormack (1969)
and Jameson and Mavriplis (1986) respectively,
employing the Cebeci and Smith (1970) model.
With this model, the pressure field generated by the
MacCormack (1969) scheme is more severe than
the one obtained with the Jameson and Mavriplis
(1986) scheme. It can be seen in the MacCormack
(1969) solution that the weaker shock before the
ramp disappears, highlighting the effect of the
turbulence model of Cebeci and Smith (1970) being
able to reduce the separation phenomenon, as seen
in Fig. 16.

8.2 Results with the Cebeci and Smith (1970)


turbulence model

Fig. 10 and Fig. 11 show the density fields


generated by the schemes of MacCormack (1969)
and Jameson and Mavriplis (1986) when the Cebeci
and Smith (1970) model is employed. Using this
model, the density field generated by the scheme of
MacCormack (1969) is denser than that generated
by the scheme of Jameson and Mavriplis (1986),
contrary to the behavior observed in the laminar
case.

156

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

scheme, with little increase in the boundary layer


thickness, and is non-existent in the Jameson and
Mavriplis (1986) solution. The weaker shock before
the ramp is non-existent in both solutions with the
Cebeci and Smith (1970) model. The circulation
bubble, originated from the boundary layer
detachment, is still perceptible in the MacCormack
(1969) solution, but is non-existent in the Jameson
and Mavriplis (1986) solution. Moreover, the
extension of the separation region detected by the
MacCormack (1969) scheme with this model is
smaller than that observed in the laminar case.

Fig. 12 Pressure contours (M-CS).

Fig. 14 Mach contours (M-CS).

Fig. 13 Pressure contours (JM-CS).

Fig. 14 and Fig. 15 show the Mach number


contours obtained by both schemes, employing the
Cebeci and Smith (1970) model. The Mach number
field generated by the scheme of Jameson and
Mavriplis (1986) is more intense than the one
generated by the MacCormack (1969) scheme, as
observed in the laminar solution.
Fig. 16 and Fig. 17 exhibit the velocity vector fields
close to the ramp wall and the respective
streamlines obtained by the schemes of
MacCormack (1969) and Jameson and Mavriplis
(1986), employing the Cebeci and Smith (1970)
model. As mentioned before, in the pressure field
analyses, the boundary layer separation is more
pronounced in the solution with the MacCormack
(1969) scheme than in the laminar solution with this

Fig. 15 Mach contours (JM-CS).

157

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

Mavriplis (1986) scheme is more severe than the


one obtained from the MacCormack (1969) scheme,
as occurred in the laminar case. A weaker shock is
again observed in the MacCormack (1969) solution,
as a result of the increase of the boundary layer
thickness caused by its separation. The Baldwin and
Lomax (1978) model does not reduce the effect of
the boundary layer detachment when the
MacCormack (1969) scheme is used, as occurred
with the Cebeci and Smith (1970) model. The
Jameson and Mavriplis (1986) scheme does not
detect this shock, having a smoother separation
region.

Fig. 16 Velocity field and streamlines (M-CS).

Fig. 18 Density contours (M-BL).

Fig. 17 Velocity field and streamlines (JM-CS).

8.3 Results with the Baldwin and Lomax (1978)


turbulence model

Fig. 18 and Fig. 19 show the density contours


obtained by adopting the schemes of MacCormack
(1969) and Jameson and Mavriplis (1986)
respectively, using the Baldwin and Lomax (1978)
model. The density field generated by the Jameson
and Mavriplis (1986) scheme is now denser than the
corresponding one generated by the MacCormack
(1969) scheme.
Fig. 20 and Fig. 21 show the pressure fields
generated by the schemes of MacCormack (1969)
and Jameson and Mavriplis (1986) respectively,
using the Baldwin and Lomax (1978) model. The
pressure field obtained from the Jameson and

Fig. 19 Density contours (JM-BL).

158

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

Fig. 20 Pressure contours (M-BL).


Fig. 23 Mach contours (JM-BL).

Fig. 22 and Fig. 23 exhibit the Mach number


contours obtained by both schemes, when applying
the Baldwin and Lomax (1978) model. The Mach
number field is more intense in the solution
generated by the Jameson and Mavriplis (1986)
scheme, as observed in the cases before.
Fig. 24 and Fig. 25 show the velocity vector fields
close to the ramp wall and the corresponding
streamlines obtained by the schemes of
MacCormack (1969) and Jameson and Mavriplis
(1986), employing the Baldwin and Lomax (1978)
model. As occurred in the laminar case, the
MacCormack (1969) scheme detects a longer
separation region and bigger increase in the
boundary layer thickness than the Jameson and
Mavriplis (1986) scheme does. It results in a weaker
shock that is formed before the ramp, as occurred in
the laminar case. The circulation bubble is also
greater in the MacCormack (1969) solution,
verifying the conjecture about the formation of the
weaker shock before the ramp. The Baldwin and
Lomax (1978) model does not minimize the
detachment effect, as in the Cebeci and Smith
(1970) model, resulting in the formation of the
weaker shock. However, with the Jameson and
Mavriplis (1986) scheme, the separation effect is
less pronounced, as in the laminar case.

Fig. 21 Pressure contours (JM-BL).

Fig. 22 Mach contours (M-BL).


159

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

Cebeci and Smith (1970) model yielded the biggest


pressure value in relation to the laminar and the
Baldwin and Lomax (1978) solutions, at x=0.325 m.
With the Jameson and Mavriplis (1986) scheme, as
in Fig. 27, there is no plateau before the ramp in the
laminar and turbulent solutions, indicating that this
scheme does not detect the weaker shock. The
largest value of pressure with this scheme occurred
when the Cebeci and Smith (1970) model was
employed, at x=0.32 m.

Fig. 24 Velocity field and streamlines (M-BL).

Fig. 26 Cp distributions (M).

Fig. 25 Velocity field and streamlines (JM-BL).

8.4 Comparison among Cp curves,


detachment and reattachment points of the
boundary layer, oblique shock angle and
simulation data

Fig. 26 and Fig. 27 highlight the Cp distributions


at the ramp wall obtained by employing the
schemes of MacCormack (1969) and Jameson and
Mavriplis (1986) respectively, involving the laminar
solution and the solutions with the two turbulence
models at the position k=kmax/2, where kmax
represents the maximum number of points in the z
direction.
In Fig. 26, solutions with the MacCormack (1969)
scheme, it has a pressure plateau at x=0.12 m in the
laminar solution, which corresponds to the weaker
shock before the ramp. The solutions with the
models of Cebeci and Smith (1970) and Baldwin
and Lomax (1978) do not have such a plateau. The

Fig. 27 Cp distributions (JM).

160

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

Fig. 28 shows the Cp distributions at the wall


obtained by using the schemes of MacCormack
(1969) and Jameson and Mavriplis (1986),
involving the laminar and turbulent solutions, at
k=kmax/2. It is noted that, in general, the Cp
distributions generated by the Jameson and
Mavriplis (1986) scheme are more severe, with
larger values of Cp than those generated by the
MacCormack (1969) scheme. However, the largest
value of Cp is obtained with the MacCormack
(1969) scheme using the Cebeci and Smith (1970)
turbulence model.
Fig. 28 Cp distributions (M and JM).
Table 1
X (m)
Detachment
Reattachment

Boundary layer detachment and reattachment points.

L (M)
0.0825
0.2100

Table 2

L (JM)
0.1125
0.1773

CS (M)
0.1275
0.1664

CS (JM)

BL (M)
0.0975
0.1937

BL (JM)
0.1200
0.1773

Shock angle obtained in the laminar and turbulent cases.

Angle (o)

L (M)

L (JM)

CS (M)

CS (JM)

BL (M)

BL (JM)

:
Error (%):

51.0

50.6

51.6

50.4

50.0

50.0

3.8

4.5

2.6

4.9

5.7

5.7

of the flow field. Anderson (1984, pp. 352 and 353)


presented a diagram with values of the shock angle,
, to oblique shock waves. The value of this angle
is determined as a function of the freestream Mach
number and of the deflection angle of the flow after
the shock wave, . Given that =20 (ramp
inclination angle) and a freestream Mach number
equals to 2.0, it is possible to obtain from this
diagram a value of equals to 53.0. Using a
transfer in figures 4, 5, 12, 13, 20 and 21,
considering the xy plane, it is possible to obtain the
values of of each scheme, in each case, as well as
the respective errors as shown in Table 2. The
smallest error is obtained from the MacCormack
(1969) scheme using the Cebeci and Smith (1970)
model.
Table 3 exhibits the numerical data of the
simulations, involving the laminar and turbulent
cases. In all simulations, a CFL number of 0.1 was
employed, which does not necessarily represent the
maximum value allowable to each scheme, in each
case.

Table 1 presents the x positions of the detachment


and reattachment of the boundary layer obtained by
the schemes of MacCormack (1969) and Jameson
and Mavriplis (1986) in laminar and turbulent cases.
As observed, the effect of the turbulence models of
Cebeci and Smith (1970) and Baldwin and Lomax
(1978) over the MacCormack (1969) scheme is a
reduced extension of the separation region in
relation to the laminar solution and hence, reducing
the effect of the increase of the boundary layer
thickness and eliminating the formation of the
weaker shock before the ramp. With the Jameson
and Mavriplis (1986) scheme, the turbulence
models practically keep, in relation to the laminar
solution, the detachment and reattachment points.
The exception is the case using the Cebeci and
Smith (1970) model, which eliminates the boundary
layer separation.
One way to quantitatively verify if the solutions
generated by each scheme are satisfactory is to
determine the shock angle of the oblique shock
wave, , measured in relation to the initial direction
161

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

Table 3

Numerical data of the simulations.

Scheme/Model
MacCormack (L)
MacCormack (CS)
MacCormack (BL)
Jameson e Mavriplis (L)
Jameson e Mavriplis (CS)
Jameson e Mavriplis (BL)

Iterations
41,515
21,670
35,132
23,473
18,815
22,773

Cost*
0.0000857
0.0001304
0.0000944
0.0002515
0.0003678
0.0002775

*Given in seconds/per cell/per iteration

The MacCormack (1969) scheme with the Baldwin


and Lomax (1978) turbulence model is the cheapest
one compared to the other implementations. The
MacCormack scheme with this turbulence model is
approximately 10.2% more expensive than the
laminar solution; however, this increase in the
computational cost is the lowest in comparison with
the other implementations. The implementations
with the Cebeci and Smith (1978) model, in both
schemes, are the most expensive ones compared to
the other implementations, being at least 45% more
expensive. This is largely a result of substantial
costing in the calculation of the boundary layer
thickness, the momentum thickness and the wall
shear stress in each iteration.

a separation region is formed at the compression


corner owing to the detachment of the boundary
layer from the wall. The results have demonstrated
that both models predict satisfactorily the
occurrence in this region, reducing, however, its
extension in relation to the laminar solution. It
agrees with the hypothesis that the turbulent
boundary layer is more stable than the laminar
boundary layer, reducing the extension of unstable
phenomenon. The Cp distributions generated by
the Jameson and Mavriplis (1986) scheme are more
severe, although the largest value of Cp is obtained
with the MacCormack (1969) scheme using the
Cebeci and Smith (1970) model. The shock angle of
the oblique shock wave is best predicted by the
MacCormack (1969) scheme using the Cebeci and
Smith (1970) model. This work is a continuation of
the study initiated by Maciel (2005a).

9. CONCLUSIONS

In the present work, the schemes of MacCormack


(1969) and Jameson and Mavriplis (1986) are
implemented by adopting the models of Cebeci and
Smith (1970) and Baldwin and Lomax (1978) to
perform turbulent flow simulations in threedimensional space. The Navier-Stokes equations in
conservative and integral forms are solved,
employing a finite volume formulation and a
structured spatial discretization. The MacCormack
(1969) scheme is a predictor/corrector method
which performs coupled time and space
discretizations of the governing equations. The
Jameson and Mavriplis (1986) algorithm is a
symmetrical scheme and its time discretization is
performed by a Runge-Kutta method of five stages.
Both schemes are second order accurate in space
and time, and require artificial dissipation to
guarantee numerical stability. Mavriplis (1990) and
Azevedo (1992) models are implemented. The
steady state physical problem of the supersonic
turbulent flow along a ramp is studied. In this study,

REFERENCES

1. Anderson JD (1984). Fundamentals of


Aerodynamics. McGraw-Hill, Inc., EUA, 563p.
2. Azevedo JLF (1992). On the Development of
Unstructured Grid Finite Volume Solvers for
High Speed Flows. Report NT-075-ASE-N,
IAE, CTA, So Jos dos Campos, SP, Brazil.
3. Baldwin BD and Lomax H (1978). Thin Layer
Approximation and Algebraic Model for
Separated Turbulent Flows. AIAA Paper
1978257.
4. Cebeci T and Smith AMO (1970). A FiniteDifference
Method
for
Calculating
Compressible Laminar and Turbulent Boundary
Layers. Journal of Basic Engineering, Trans.
ASME, Series B, 92(3):523535.
5. Fox RW and McDonald AT (1988). Introduo
Mecnica dos Fluidos. Ed. Guanabara
Koogan, Rio de Janeiro, RJ, Brazil, 632p.
162

Engineering Applications of Computational Fluid Mechanics Vol. 1, No. 3 (2007)

15. Maciel ESG (2006c). Estudo de Escoamentos


Turbulentos Utilizando os Modelos de Cebeci e
Smith e de Baldwin e Lomax e Comparao
entre os Algoritmos de MacCormack e de
Jameson e Mavriplis. Proceedings of the 7th
Symposium of Computational Mechanics (VII
SIMMEC). Arax, MG, Brazil.
16. Maciel ESG (2006d). Comparao entre
Diferentes Modelos de Dissipao Artificial
Aplicados a um Sistema de Coordenadas
Generalizadas Parte I. Proceedings of the 7th
Symposium of Computational Mechanics (VII
SIMMEC). Arax, MG, Brazil.
17. Mavriplis DJ (1990). Accurate Multigrid
Solution of the Euler Equations on
Unstructutred and Adaptive Meshes. AIAA
Journal 28(2):213221.
18. Swanson RC and Radespiel R (1991). Cell
Centered and Cell Vertex Multigrid Schemes
for the Navier-Stokes Equations. AIAA Journal
29(5):697703.

6. Jameson A and Mavriplis DJ (1986). Finite


Volume Solution of the Two-Dimensional
Euler Equations on a Regular Triangular Mesh.
AIAA Journal 24(4): 611618.
7. Jameson A, Schmidt W and Turkel E (1981).
Numerical Solution for the Euler Equations by
Finite Volume Methods Using Runge-Kutta
Time Stepping Schemes. AIAA Paper
19811259.
8. MacCormack RW (1969). The Effect of
Viscosity in Hypervelocity Impact Cratering.
AIAA Paper 1969354.
9. Maciel ESG (2002). Simulao Numrica de
Escoamentos Supersnicos e Hipersnicos
Utilizando Tcnicas de Dinmica dos Fluidos
Computacional. Doctoral Thesis, ITA, So Jos
dos Campos, SP, Brazil, 258p.
10. Maciel ESG (2005a). Comparao entre os
Algoritmos de MacCormack e de Jameson e
Mavriplis na Soluo das Equaes de Euler e
de Navier-Stokes no Espao Tridimensional.
Mecnica
Computacional
Journal
XXIV(12):20552074.
11. Maciel ESG (2005b). Solution of the NavierStokes Equations in the Three-Dimensional
Space Using the MacCormack Algorithm Part
II. Proceedings of the XVIII International
Congress of Mechanical Engineering (XVIII
COBEM). Ouro Preto, MG, Brazil.
12. Maciel ESG (2005c). Solution of the NavierStokes Equations in the Three-Dimensional
Space Using the Jameson and Mavriplis
Algorithm Part II. Proceedings of the XVIII
International
Congress
of
Mechanical
Engineering (XVIII COBEM). Ouro Preto, MG,
Brazil.
13. Maciel ESG (2006a). Comparao entre os
Modelos de Turbulncia de Cebeci e Smith e de
Baldwin e Lomax. Proceedings of the 5th
Spring School of Transition and Turbulence (V
EPTT). Rio de Janeiro, RJ, Brazil.
14. Maciel ESG (2006b). Relatrio ao Conselho
Nacional de Pesquisa e Desenvolvimento
Tecnolgico (CNPq) sobre as Atividades de
Pesquisa Desenvolvidas no Terceiro Ano de
Vigncia da Bolsa de Estudos para Nvel DCRIF Referente ao Processo n 304318/2003-5.
Report, National Council of Research and
Technological Development (CNPq). Brazil,
August, 52p.

163

S-ar putea să vă placă și