Vous êtes sur la page 1sur 114

Helsinki University of Technology. Laboratory of Aerodynamics.

Series B
Teknillinen korkeakoulu. Aerodynamiikan laboratorio.
Sarja B
Espoo 2001, FINLAND

IMPLEMENTING AN EXPLICIT ALGEBRAIC REYNOLDS STRESS


MODEL INTO THE THREE-DIMENSIONAL FINFLO FLOW SOLVER

Report B-52

Ville Hmlinen

TEKNILLINEN KORKEAKOULU
TEKNISKA HGSKOLAN
HELSINKI UNIVERSITY OF TECHNOLOGY

Helsinki University of Technology. Laboratory of Aerodynamics.


Series B
Teknillinen korkeakoulu. Aerodynamiikan laboratorio.
Sarja B
Espoo 2001, FINLAND

IMPLEMENTING AN EXPLICIT ALGEBRAIC REYNOLDS STRESS


MODEL INTO THE THREE-DIMENSIONAL FINFLO FLOW SOLVER

Report B-52

Ville Hmlinen

Approved by Seppo Laine

Helsinki University of Technology


Department of Mechanical Engineering
Laboratory of Aerodynamics
Teknillinen korkeakoulu
Konetekniikan osasto
Aerodynamiikan laboratorio

Distribution:
Helsinki University of Technology
Laboratory of Aerodynamics
P.O.Box 4400
FIN-02015 HUT
Tel. +358-9-451 3421
Fax +358-9-451 3418
Ville Hmlinen
ISBN 951-22-5605-3
ISBN 951-22-6824-8 (PDF)
ISSN 1456-6990
Printed in Otamedia
Espoo 2001, FINLAND

Abstract

This thesis describes the implementing of an explicit algebraic Reynolds


stress model (EARSM) into the three-dimensional FINFLO flow solver. The
EARSM replaces the Boussinesq eddy-viscosity assumption by a more
general constitutive relation for the second-order correlation in the Reynolds averaged Navier-Stokes equations.
The thesis begins by describing the Navier-Stokes equations, in both their
time-accurate and Reynolds averaged form. Some general aspects of turbulence are also discussed and a differential Reynolds stress model is presented. Then an algebraic Reynolds stress model (ARSM) is formulated
and simplified to its explicit form. After that some programming aspects of
the model are discussed. Finally, four test cases are selected, calculated
and analysed to verify the correct implementation of the model.
As the first test case a two-dimensional boundary layer over a flat plate is
calculated to verify the connection between the EARSM and the rest of the
FINFLO. The second order effects of the model are studied with the second
test case, a fully developed flow inside a rectangular duct. Thirdly, a fully
developed flow inside a rotating pipe is calculated to examine the third order effects of the model. As the last test case an axial flow over a cylinder,
part of which rotates, is calculated to further study the behaviour of the
EARSM when a sudden strain component is applied to the flow field. From
the results of the test cases it is then concluded that the EARSM has been
correctly implemented. The model gives significant improvements in most
cases compared to standard eddy-viscosity models without a large increase in the computational effort.

iii

Contents

Abstract .................................................................................................... iii


Contents .................................................................................................... v
Nomenclature.......................................................................................... vii
1

Introduction......................................................................................1

Governing equations and eddy-viscosity turbulence models ......3


2.1

Navier-Stokes equations ........................................................3

2.2

Reynolds averaging................................................................5

2.3

General facts of turbulence.....................................................6

2.4

Eddy-viscosity turbulence models...........................................8


2.4.1 Algebraic models.........................................................9
2.4.2 One-equation models................................................10
2.4.3 Two-equation models................................................11

Differential Reynolds stress models ............................................15


3.1

Reasons for using the Reynolds stress models ....................15

3.2

Exact equation for the Reynolds stress tensor......................16

3.3

Modelling the Reynolds stress tensor ...................................18

Formulation of an explicit algebraic Reynolds stress model .....21


4.1

Algebraic Reynolds stress model..........................................21

4.2

Explicit algebraic Reynolds stress model..............................26

4.3

Simplified EARSM for two-dimensional mean flow................29

4.4

Simplified EARSM for three-dimensional mean flow .............31

4.5

Solution of the general quasi-linear ARSM equation.............33

Programming approach and written subroutines........................37


5.1

Subroutine EARSM1 ............................................................37

5.2

Subroutine TTIMES ..............................................................41

5.3

Subroutine ANISFLUX..........................................................42

5.4

Subroutine FLUXP................................................................43

5.4

Subroutine PEEKOO ............................................................43

5.5

Subroutine VELGRAD ..........................................................44

5.6

Boundary routines ................................................................44

Contents

Test cases ......................................................................................47


6.1

Flow over a flat plate.............................................................48


6.1.1 Grid ...........................................................................48
6.1.2 Boundary conditions..................................................50
6.1.3 Results ......................................................................51

6.2

Flow inside a rectangular duct ..............................................55


6.2.1 Grid ...........................................................................55
6.2.2 Boundary conditions..................................................56
6.2.3 Results ......................................................................57

6.3

Flow inside a rotating pipe ....................................................63


6.3.1 Grid ...........................................................................64
6.3.2 Boundary conditions..................................................65
6.3.3 Results ......................................................................65

6.4

Flow over a rotating cylinder .................................................69


6.4.1 Grid ...........................................................................70
6.4.2 Boundary conditions..................................................71
6.4.3 Results ......................................................................71

Summary and conclusions............................................................79

Bibliography ............................................................................................81
Appendix A, Subroutine EARSM1..........................................................87
Appendix B, Subroutine TTIMES ...........................................................91
Appendix C, Subroutine ANISFLUX.......................................................93
Appendix D, Subroutine FLUXP.............................................................95
Appendix E, Subroutine PEEKOO .........................................................97
Appendix F, Subroutine VELGRAD........................................................99

vi

Nomenclature

Roman alphabet

Model coefficient

aij
C
Cf
Cij( a )
C eff
E
e
fi
H
J
k
l
N
ni
P
Pij
p
qi
Re
S
S ij
T
t
ui
u
u

Reynolds stress anisotropy tensor

Cell volume

xi

Model coefficient
Skin friction coefficient
Coriolis tensor
Effective eddy-viscosity coefficient
Total energy
Specific internal energy
Body force vector
Matrix defined by the equation (4.44)
Matrix defined by the equation (4.45)
Turbulent kinetic energy, Heat transfer coefficient
Turbulent length scale
Symbol defined by the equation (4.25), Non-dimensional rotation rate
Cell face unit normal vector
Production of turbulence
Production of turbulence tensor
Pressure
Heat flux vector
Reynolds number
Cell face area
Mean strain tensor
Temperature
Time
Velocity vector
Friction velocity
Fluctuating component of the u
Cartesian coordinate vector
Dimensionless wall distance

vii

Nomenclature

Greek alphabet

ij

ij
ijk
ij

w
ij
T
ij
*ij
ijR
ijS

kS

Coefficient for the terms of the aij


Kroneckers delta
Dissipation rate
Dissipation rate tensor
Permutation tensor
Redistribution tensor
Molecular viscosity
Density
Turbulent time-scale
Wall shear stress
Viscous stress tensor
Kinematic eddy-viscosity
Mean vorticity tensor
Absolute mean vorticity tensor
Effective mean vorticity tensor
System rotation tensor

k specific dissipation rate


Constant system rotation rate vector

Invariants

II S

Second invariant of the S ij

II

Second invariant of the ij

III

Third invariant of the S ij

IV

Third invariant defined as S ij jk ki

Fourth invariant defined as S ij S jk kl li

viii

Chapter 1
Introduction

In earlier days fluid dynamics, like other physical sciences, was divided into
theoretical and experimental branches. The equipment and vehicles involving fluid flow were designed and analysed by these two methods. With
the evolution of the digital computer, a third method called Computational
Fluid Dynamics (CFD) has become available. In this computational approach the equations that govern a process of interest are solved numerically at certain discrete points of space.
The evolution of numerical methods for solving ordinary and partial differential equations began approximately at the beginning of the twentieth
century. The automatic digital computer was invented in the early 1940s
and was used from nearly the beginning to solve problems in fluid dynamics. The explosion in computational activity did not, however, begin until the
high-speed digital computers began to be generally available in the 1960s
(Ref. [1]).
The three key elements of CFD are algorithm development, grid generation
and turbulence modelling. In this study only the third element is scrutinised.
Turbulence is inherently three-dimensional and time dependent, and an
enormous amount of information is thus required to completely describe a

Introduction

turbulent flow. This is beyond the capability of the existing computers for
virtually all practical flows. Thus, some kind of approximate and statistical
method, called a turbulence model, is needed.
Helsinki University of Technology has quite a long history in the field of
CFD. The development of the parallel multi-block Navier-Stokes flow solver
called FINFLO was initiated in 1987. The original two-man development
group has grown along with new applications. Nowadays, in 1995 established CFD group consists of about 15 researchers from the Laboratory of
Aerodynamics, the Laboratory of Applied Thermodynamics and the Laboratory of Ship Hydrodynamics, all utilising the FINFLO code. The FINFLO is
nowadays suitable for compressible, time dependent, laminar and turbulent
flows and has been applied to dozens of demanding research and development projects. The code is able to handle structured multi-block grids
and the equations are solved by an implicit pseudo-time integration scheme
using Roes flux splitting.
The CFD-group has always tried to keep up with the most recent turbulence modelling. Prior to this work the FINFLO included the following turbulence models; Baldwin-Lomax [2], Cebeci-Smith [3], Chiens low Reynolds number k model [4] and different variants of Menters k
model (BSL), (SST) [5] [6] and (RCSST) [7]. The aim of this thesis has
been to implement into the FINFLO a new turbulence model called the Explicit Algebraic Reynolds Stress Model (EARSM) developed by Stefan Wallin and Arne V. Johansson from Sweden. The formulation of the model is
presented briefly in this thesis, but the reader is encouraged to find the
complete description in reference [8]. The EARSM was programmed using
Fortran 77 and 90 as they are used throughout the main code. After the
programming was completed, four test cases were selected, calculated and
analysed to verify the correct implementation of the model.

Chapter 2
Governing equations and eddy-viscosity turbulence models

In this chapter the equations governing the flow field are presented in their
time-accurate and Reynolds averaged form. Then some general remarks of
turbulence are made and importance for the turbulence modelling is emphasised. At the end of the chapter the basic aspects of the so-called eddyviscosity turbulence models are discussed.

2.1

Navier-Stokes equations

The equations of viscous flow have been known for more than 100 years.
The exact number of basic equations depends upon personal preference,
but some relations are, however, more basic than others. Usually the basic
system of equations is considered to be the three laws of conservation for
physical systems:
- Conservation of mass (i.e. the continuity equation)
- Conservation of momentum (i.e. the Newtons second law)
- Conservation of energy (i.e. the first law of thermodynamics)

Governing equations and eddy-viscosity turbulence models

The continuity equation simply states that the mass must be conserved. In
the Cartesian coordinates xi this equation can be written as

( u i )
+
=0
t
xi

(2.1)

where is the density of the fluid, t time and u i the velocity vector. Tensor notation will be used throughout the text whenever practical. The second equation, conservation of momentum, states that momentum must be
conserved. It can be written in the Cartesian coordinates as

( u i ) (u i u j )
p ij
+
= f i
+
t
xi
xi x j

(2.2)

where f i is a body force, p the pressure. The viscous stress tensor ij is

u u 2 u
ij = i + j ij k
x j xi 3 xk

(2.3)

where is the molecular viscosity and ij the Kroneckers delta. The third
equation, conservation of energy, states that the energy must be conserved. It can be written as

+
[u j (E + p )] (ui ij q j ) = 0
t x j
x j

(2.4)

where E is defined as the total energy and can be written as

E = e + u i ui
2

(2.5)

where e is the specific internal energy. Assuming Fouriers heat transfer


law the heat flux vector can be written as

qi = k

T
xi

where k in the heat transfer coefficient and T the temperature.

(2.6)

Governing equations and eddy-viscosity turbulence models

2.2

Reynolds averaging

In a turbulent flow the local pressure, density, velocity components and


temperature vary randomly with time. A reasonable approach is to separate
the flow quantities into stationary and random parts. The quantities are thus
usually presented as a sum of the mean flow value and the fluctuating part
as u i + u i . This formulation is inserted into the equations (2.1), (2.2) and
(2.4); a process called Reynolds averaging is performed. There are three
forms of averaging present in the turbulence-model research, which are a
time average, a spatial average and an ensemble average. The general
term used to describe these averaging processes is mean. The most
commonly used averaging is the first, time averaging. It can be used only
for stationary turbulence, i.e. for a turbulent flow that does not vary with
time on the average. For such a flow the mean flow value is defined as

1
ui = lim
T T

t +T

u (t )dt

(2.7)

In practice taking the limit to infinity means that the integration time T
needs to be long enough relative to the maximum period of the assumed
velocity fluctuations.
The Reynolds averaged Navier-Stokes equations (RANS) are obtained
from the continuity and momentum equations, (2.1) and (2.2), by taking the
time average of all the terms in the equations. The continuity equation does
not change since it is linear in terms of the velocity. However, the momentum equation is non-linear, which means that all the fluctuating components
do not vanish. An extra term, called Reynolds stress u iu j , appears in the
momentum equation (2.2). The result can be written as

( ui ) (ui u j )
p ij uiu j
+
= f i
+
t
xi
xi
x j

(2.8)

where the overbar has been dropped from the mean values. This convention will be used throughout the text whenever obvious. The equation (2.8)
presents the fundamental problem of turbulence. In order to compute all the

Governing equations and eddy-viscosity turbulence models

mean-flow properties of the turbulent flow we need a reasonably accurate


way to compute the Reynolds stress u iu j . This is the fundamental reason
for the need of the turbulence models.
The complete set of the Reynolds averaged Navier-Stokes equations are
not presented in this text, because they can be found from many references, for example [9]. Scalar transport equations are also needed, for example to describe the transport of the concentration of species or the mass
fraction of species. Their exact formulation can be found from reference [9]
or [10]. The presentation of the exact formulation of the FINFLO along with
the solution methods used is beyond the scope of this text. The reader is
encouraged to consult references [11], [12] and [13].

2.3

General facts of turbulence

Many basic definitions of turbulence have been written, but probably the
most accurate was formed by Bradshaw in 1974:

"Turbulent fluid motion is irregular condition of flow in which the various


quantities show a random variation with time and space co-ordinates, so
that statistically distinct average values can be discerned. Turbulence has a
wide range of scales." (Ref. [14])
Analysis of the solutions to the Navier-Stokes equations shows that turbulence develops as an instability of the laminar flow. At low Reynolds numbers, fluid layers slide smoothly past each other and the molecular viscosity
dampens the randomly occurring small-scale, high frequency disturbances.
The flow is laminar and exactly predictable. When Reynolds number increases, the restraining effects of viscosity are too weak to prevent such
disturbances from amplifying. The disturbances grow gradually and become non-linear. They also interact with neighbouring disturbances. At high
Reynolds number, the flow reaches a chaotic and non-repeating, i.e. turbulent, state. Turbulent motions are characterised by the following properties: three dimensional, unsteady, random, strongly vorticial, strongly dissipative and strongly diffusive. Turbulence increases friction, heat transfer

Governing equations and eddy-viscosity turbulence models

and mixing and spreading rate. It also reduces the separation tendency of
the flow by energising the boundary layer (Ref. [15]).
The non-linearity of the Navier-Stokes equations leads to interactions between turbulent fluctuations of different wavelengths and directions. According to reference [14] the wavelengths of motion usually extend all the
way from a maximum comparable to the width of the flow to a minimum
fixed by viscous dissipation of energy. The main physical process that
spreads the motion over a wide range of wavelengths is called vortex
stretching. The turbulence gains energy if the vortex elements are primarily
oriented in a direction in which the mean velocity gradients can stretch
them. This is called production of turbulence. The turbulent kinetic energy is
then convected, diffused and dissipated.
The larger-scale turbulent motion carries most of the energy and is mainly
responsible for the enhanced diffusivity and attending stresses. The large
eddies have memory and their orientation is sensitive to the mean flow.
The large eddies randomly stretch the smaller eddies, cascading energy to
them. Small eddies lose their orientation preference and become statistically isotropic. Energy is dissipated by viscosity in the shortest wavelengths, although the long-wavelength motion at the start of the cascade
sets the rate of dissipation of energy. The shortest wavelengths simply adjust accordingly. This kind of turbulence is called equilibrium turbulence.
The ratio of largest to smallest scales increases rapidly as the Reynolds
number increases (Ref. [14]).
If the unsteady Navier-Stokes equations were calculated, a vast range of
time and length scales would have to be computed. In other words, a very
fine grid and very small time steps would be required. This is possible with
todays computers for only very simple problems with small Reynolds numbers. Thus a turbulence model is needed to account at least some of the
fluctuating motion by a statistical approach. The model should be, in principle, simple, broadly general and applicable, based on physics rather than
intuition, computationally stable and coordinately invariable. This is, of
course, an impossible list of requirements. The next sections of this chapter
describe traditional ways to model the turbulence using a so-called Boussinesq eddy-viscosity approximation.

Governing equations and eddy-viscosity turbulence models

2.4

Eddy-viscosity turbulence models

In this section the so-called eddy-viscosity turbulence models are briefly


discussed. The models can be divided in three categories; zero-, one- and
two-equation models. These models use the Boussinesq eddy-viscosity
approximation. The word approximation is a little misleading here because
it is really more of a hypothesis than an approximation. It can be written for
one- and two-equation models as

u j u i 2 u k 2
ij
u iu j = T
+
ij k
xi x j 3 xk 3

(2.9)

For the zero-equation models, which do not make use of the turbulent kinetic energy k u iui 2 , the last term of the equation is zero. The Boussinesq approximation is used to compute the Reynolds stress tensor as the
product of the kinematic eddy-viscosity T and the mean strain-rate tensor.
The kinematic eddy-viscosity will be referred hereafter as the eddy-viscosity
for simplicity. The Reynolds stress components are calculated from the
eddy-viscosity for an incompressible flow as

2
uiu j = 2 T S ij ij k
3

(2.10)

where the mean strain-rate tensor is defined as

1 u u j
S ij = i +
2 x j xi

(2.11)

The eddy-viscosity is a scalar field variable and the different Reynolds


stress components of a certain point of the flow field thus scale with the
mean strain-rate tensor, i.e. the Reynolds stress components are linearly
proportional to the mean strain-rate tensor. Different ways to calculate the
eddy-viscosity are described in the next subsections.
The equations (2.10) and (2.11) as well as all the following equations in this
text are presented for an incompressible flow. The FINFLO calculation routine calculates the flow as compressible, of course, while all of the turbu-

Governing equations and eddy-viscosity turbulence models

lence models approximate the flow as incompressible. This approximation


is a so-called Morkovins hypothesis meaning that the compressibility affects the turbulence quantities only at hypersonic speeds.

2.4.1 Algebraic models


Models where the eddy-viscosity is completely determined in terms of the
local mean flow variables are referred to as zero-equation or algebraic
models. Algebraic models are the simplest of all turbulence models. They
often use some form of Prandtls mixing length hypothesis in order to compute the eddy-viscosity using terms of mixing length in analogy to the molecular mixing phenomenon. Whereas the molecular viscosity is an intrinsic
property of the fluid, the eddy-viscosity (and hence the mixing length) depends upon the flow. Because of this, the eddy-viscosity and mixing length
must be specified in advance by an algebraic relation between eddyviscosity and length scales of the mean flow. Thus, algebraic models are,
by definition, incomplete models of turbulence. (Ref. [14]).
The eddy-viscosity is usually calculated as

2
T = l mix

dui
dy j

(2.12)

where l mix is the mixing length analogous with the molecular mean free
path in molecular mixing.
Well-known and much used algebraic models are the Baldwin-Lomax [2]
and the Cebeci-Smith [3] models, both of which are two-layer models with
separate expressions for the computation of the eddy-viscosity in each
layer. The Cebeci-Smith model is simple and easy to implement. Most
computational effort goes to the calculation of the boundary layer velocity
thickness. The Baldwin-Lomax model was formulated for use in computations where boundary layer properties such as the boundary layer and the
boundary layer velocity thickness are difficult to determine.
Algebraic models are conceptually very simple, economical in terms of
computer resources and rarely cause any unexpected numerical difficulties.

Governing equations and eddy-viscosity turbulence models

The user of the algebraic models must always be aware of the issue of incompleteness. Both the Cebeci-Smith and Baldwin-Lomax models work
well only for the boundary layer flows for which they have been fine-tuned.
They reproduce skin friction and velocity profiles well for incompressible
turbulent boundary layers provided the pressure gradient is not too strong.
However, they can not be applied to other flow cases, such as wake flows
or free jets.

2.4.2 One-equation models


In one-equation models, one of the two turbulence scales or a combination
of both is determined from a transport equation. The transport equation for
the turbulence kinetic energy can be written as

Dk
ui

+
= ij
Dt
x j
x j

k 1

1
x 2 uiuiu j pu j
j

(2.13)

where the tensor notation D Dt t + u j x j is used to denote the rate


of change following the mean flow. The term ij ui x j is the production
of the turbulence, P , the term k x j is the molecular diffusion, the term

uiuiu j 2 is the turbulent flux of the turbulent kinetic energy, k , and the last
term p u j is the so-called pressure diffusion term. The last term is usually neglected because its contribution is very small.
The is the dissipation rate per unit mass, defined as

u i u i
xk xk

(2.14)

The eddy-viscosity is usually calculated as

T = C l mix k

(2.15)

where the C is a model coefficient. As for algebraic models, additional


information from the mean flow field or geometrical measures is needed.
Most commonly used one-equation models are the Spalart-Allmaras [16]

10

Governing equations and eddy-viscosity turbulence models

and Baldwin-Barth [17] models. Actually, in the models the transported


quantity is k 2 , which follows from the definition of the turbulent Reynolds
number ReT uT lT , since uT ~ k 1 2 and lT ~ k 3 2 .
According to reference [15], the one-equation models give usually more
accurate results than algebraic models because more realism is obtained
as a result of deducing the turbulent velocity scale from a turbulence quantity

k rather than from gradients of mean velocity. This also implies that,

by obtaining the solution for the turbulent velocity scale from a differential
transport equation, some account is taken of the history effects, which become important in non-equilibrium flows. The differential transport equation
also removes the problems associated to the flow field points where the
local mean velocity gradient is zero.
The main drawbacks with one-equation models are largely the same as for
the mixing-length models. The history effects are not accounted for the
length scale l, which is still prescribed as an algebraic function of local
quantities. According to reference [18] the Spalart-Allmaras model performs
better than the Chiens k model in a decelerating flow with adverse
pressure gradient because it is specially designed for aerodynamical purposes. It is one of the most popular turbulence models in the field of aeronautical applications, especially in the U.S.A.

2.4.3 Two-equation models


Two-equation models have served as the foundation for much of the turbulence model research until very recent years. The most significant difference between the other eddy-viscosity models and the two-equation models is that the latter are complete, i.e. can be used to predict properties of a
given turbulent flow with no prior knowledge of the turbulence structure or
flow geometry. These models provide an equation not only for computation
of k , but also for the turbulence length scale or equivalent.
The starting point for virtually all linear two-equation eddy-viscosity models
is the Boussinesq approximation (2.9) and the turbulence kinetic energy
equation (2.13). The choice of the second variable is arbitrary and many

11

Governing equations and eddy-viscosity turbulence models

proposals have been presented. By far two of the most popular dependent
variables for the second variable have been the dissipation rate and the

k -specific dissipation rate . Nowadays there is also increasing interest in


the use of the turbulent time-scale . The most common proposals are
presented in the table 2.1
Table 2.1. Proposals for the second variable (Ref. [15]).
Proposer

Year

Variable

Symbol

Physical meaning

Kolmogorov

1942

k 1 2 l 1

Rotta

1951

Length scale

Rotta

1968

kl

kl

Harlow & Nakayama

1968

k 3 2 l 1

Energy dissipation rate

Spalding

1969

kl 2

Vorticity fluctuations squared

Nee & Kovasznay

1969

lk 1 2

Eddy viscosity

Speziale

1992

lk 1 2

Frequency

times the length scale

Time-scale

The k model is the best-known two-equation turbulence model because it is simple to understand and use and relatively easy to program.
The model dates back to the late 1960s but the most used formulation, referred as the standard k model, was presented in 1972 by Jones and
Launder [19]. The eddy-viscosity used with the Boussinesq approximation
is calculated as

T = C

k2

(2.16)

where C is a model coefficient. The k model is used widely in practical engineering calculations even though the standard model can be used
only with attached flows with thin shear layers. The model fails to reproduce
the correct flow behaviour in many important flow situations, such as: ad-

12

Governing equations and eddy-viscosity turbulence models

verse pressure gradients, bluff-body flows, separation, streamline curvature, swirl, buoyancy, turbulence driven secondary motion, compressibility
and unsteadiness.
The k model is based on the choice of the specific dissipation rate as
the second variable. The eddy viscosity is calculated as

T =

(2.17)

Even though the wall boundary condition for the is more difficult to formulate and program than for the , the k model has found its way to
many engineering, especially aerodynamic, applications. This is mainly due
to the fact that it can also reproduce the flow field behaviour with adverse
pressure gradient. It usually also predicts separation better than the other
linear two-equation eddy-viscosity models. Actually, the k SST model
[7] predicts the separation fairly well, but it is no longer really a linear
model.

13

Chapter 3
Differential Reynolds stress models

In this chapter a different approach to model the turbulence is reviewed.


The Boussinesq approximation is abandoned and the transport equation for
the Reynolds stresses is presented instead. In the end of the chapter modelling propositions of the different terms of the transport equation are discussed.

3.1

Reasons for using the Reynolds stress models

Eddy-viscosity models perform reasonably well in attached boundary layer


flows as long as only one component of the Reynolds stress tensor is of
significant importance. In these cases the eddy-viscosity can be thought as
a representative for the significant Reynolds stress component. However, if
the flow becomes more complicated the eddy-viscosity assumption fails.
There is thus not much hope for a more general validity of the eddyviscosity approach (Ref. [8]).
The development of the so-called differential Reynolds stress models (also
called Reynolds stress transport, second-order closure, secondmoment closure and second-order modelling) started officially in 1968,

15

Differential Reynolds stress models

when C. Donaldson lectured in Stanford of the need to model the Reynolds


stress terms. The Boussinesq eddy-viscosity hypothesis is abandoned and
the unknown Reynolds stress components are obtained directly from the
solution of differential transport equations in which they are the dependent
variables. The Reynolds stress tensor is symmetric meaning that four
equations need to be solved in two-dimensional and six equations in threedimensional flows. An additional equation for the length scale or equivalent
is also solved in each case.
The Reynolds stress models are thus more complicated than the eddyviscosity models. Still, they are attractive because they provide a more accurate representation of the turbulence and are valid over a wider range of
flows. They can capture many of the complex effects encountered in nature
and in engineering practice without recourse to the ad-hoc modifications
found necessary in lower-order models. According to reference [15], examples of flow situations where Reynolds stress models were found to give
accurate predictions include flows with streamline curvature, rotation, swirl
and buoyancy.

3.2

Exact equation for the Reynolds stress tensor

The exact transport equation for the Reynolds-stress tensor u iu j is obtained from the momentum equation (2.2) by multiplying the instantaneous

u i component equation by u j , the u j component equation by ui , adding


the two and then time-averaging the result. Following reference [20], the
result can be written for constant-density flows with no body forces after
some rearrangement as

u j

u
Du iu j
= uiu k
Terms I and II
+ u j u k i
t
xk
xk

u iu j

1

pui jk + pu j ik
Term III
uiu j u k +

xk
xk

p u i u j
+
x j xi

u j

2 ui

xk xk

Terms IV and V
(3.1)

16

Differential Reynolds stress models

Term I is the convection term. It represents the rate of change of u iu j


along a streamline. In steady flows this is equal to the rate at which Reynolds stresses are convected by the mean fluid motion.
Term II is the production term. It represents the rate of production of u iu j
by mean shear, which is the reason for turbulence as discussed earlier.
The shear stress is generated by interaction of transverse normal stress
and shear strain. The production term is both large and exact, which is a
partial reason for the success of the Reynolds stress models. It will be
hereafter referred to as Pij .
Term III is the diffusion term. It represents the rate of spatial transport of

u iu j by the action of turbulent fluctuations, pressure fluctuations and molecular diffusion.


Term IV is the redistribution term, which is also called the pressure-strain
term. It represents the redistribution of the available turbulent kinetic energy
amongst the fluctuating velocity components. The mean flow direction being x1 , only u1u1 is generated in the shear layer, meaning that it is usually
much larger than the other Reynolds stress components. The redistribution
term shares u1u1 s energy out to u 2 u 2 and u 3 u 3 . It has no effect on the
overall level of turbulent kinetic energy since it has zero trace. The redistribution term just drives turbulence towards isotropy by redistributing energy.
This term will be hereafter referred to as ij .
Usually u1u 2 has the opposite sign to the boundary layer shear strain. The
redistribution term 12 reduces also u1u 2 since isotropic turbulence must
be shear-free. The same applies to the possible shear strains to other directions.
Term V is the dissipation term. It represents the dissipation rate of u iu j
due to molecular viscous action. The dissipation term will be hereafter referred to as ij .

17

Differential Reynolds stress models

3.3

Modelling the Reynolds stress tensor

With the exception of the convection and production terms of the equation
for the Reynolds stress tensor, all the other terms introduce new unknown
correlations that must be modelled in terms of known or knowable quantities in order to close the equations.
The diffusion term is a sum of three parts, which are called the viscous,
pressure and turbulent diffusion. The contribution of the viscous diffusion
term to the total rate of transport of u iu j is small at high values of the turbulence Reynolds number. This term is thus often neglected in high Reynolds number applications even though it could be included in the numerical simulations without any difficulty.
Little is known about the pressure diffusion term since direct measurements
can not be conducted. Estimates of its magnitude are done by indirect
methods, which contain all the errors made in the measurements of the
other terms. The consensus of several experiments, however, suggests
that this term is relatively unimportant and may therefore be neglected (Ref.
[15]).
Daly and Harlow [21] were the first to model the turbulent diffusion term
using the gradient transport hypothesis, meaning that the diffusion of a
quantity is assumed to be proportional to the spatial gradient of the same
quantity. According to reference [15] the turbulent diffusion can be written
as

uiu j u k = C S

uiu j
k
u k ul

xl

(3.2)

where C S is an empirical model coefficient typically set equal to 0.22,


which is obtained by computer optimisation. The ratio k represents a
characteristic time scale of the energy containing eddies.
Alternative models for the turbulent diffusion exist, but they all are considerably more complex than Daly & Harlows model without necessarily pro-

18

Differential Reynolds stress models

ducing better overall performance. Examples of such models are Hanjalic &
Launder [22] and Lumley & Khajeh-Nouri [23].
The redistribution term modelling is based on the Poisson equation for
the instantaneous pressure. It is obtained by differentiating the NavierStokes equations in a manner that is beyond the scope of this text. Then,
by subtracting the mean, the following equation for the fluctuating pressure
is obtained

2 uiu j uiu j
ui ui
1 2 p
=

+
2

xi2
x j xi
x j x j

(3.3)

where a sum of two very distinct terms is identified in the parentheses. The
first term of the sum contains only turbulence quantities and second term
contains mean-velocity gradients. When the equation is solved for homogeneous turbulence, the so-called Chous integral for the pressure-strain
correlation is obtained

1
p ui
=
x j 4
where

3 ul*u m* ui
ul 2 u m* ui dvol
2
+

rl rm rj xm rl r j r + ij , w

(3.4)

denotes quantities evaluated at position x * and r = x * x . ij,w

is a surface integral, which is important only when the typical size of the
energy containing eddies is of the same order as distance from a wall (Ref.
[15]).
Launder, Reece and Rodi (LRR) proposed that the first and the second
term in the parentheses of the equation (3.4) are modelled separately [24].
This is the traditional approach to modelling Chous integral and the great
majority of Reynolds-stress model developers have followed it.
Speziale, Sarkar and Gatski (SSG) proposed that the Chous integral could
be modelled as a whole [25]. This approach is rapidly becoming more
popular than (LLR) as it does not appear to require a specific model for the
surface integral ij,w .

19

Differential Reynolds stress models

The dissipation term is based on the fact that the molecular viscosity converts the turbulent kinetic energy into internal heat by acting on the smallscale, high frequency motions. Normally these motions can be assumed to
be isotropic. This implies that the dissipation rate of the Reynolds stress
component may be related to , the total dissipation rate of the turbulent
kinetic energy as

2
ij = ij
3

(3.5)

First the total dissipation rate is modelled through a transport equation.


The equation is similar to the second equation of the k models. Certainly, other alternatives to exist, such as or . The equation for the

can be written as
D
k

=
C u k ul
+ (C 1 Pkl S lk C 2 )
Dt xk
xl k

(3.6)

where C , C 1 and C 2 are model coefficients and Pkl the production of


the stress component S kl . The diffusion of the dissipation rate can now be
modelled using the generalised gradient diffusion model by Daly & Harlow.
Further details are beyond the scope of this text.
The final task before the differential Reynolds stress model can be implemented is the determination of the model coefficients, because the modelling approximations involve a number of empirical coefficients for which
values must be determined to close the Reynolds stress equations. Usually
six basic criteria are used for determining those coefficients, which are
mentioned briefly here. First, grid-generated turbulence must decay with a
certain experimentally observed rate in the absence of shear. Secondly,
turbulence must adjust when a rate of strain is suddenly applied to an isotropic turbulence field. Thirdly, homogenous free shear data can be used to
match the production to dissipation. The model must also be able to predict
the boundary layer close to a solid wall and turbulence must collapse in
stabilising curvature. Finally, a computer can be used to optimise the remainder of the coefficients using large number of well-documented simple
shear layers (Ref. [15]).

20

Chapter 4
Formulation of an explicit algebraic Reynolds
stress model

In this chapter the explicit algebraic Reynolds stress model (EARSM) is


formulated. First the exact differential Reynolds stress model is approximated with an algebraic Reynolds stress model (ARSM). The algebraic
Reynolds stress model is then modified into an explicit form. The simplified
solutions are presented for two- and three-dimensional mean flow. At the
end of the chapter the non-simplified solution for a three-dimensional mean
flow is also presented. This chapter, as well as the complete EARSM, is
based on reference [8]. There are also, of course, other formulations of the
EARSM, such as the one presented in reference [26].

4.1

Algebraic Reynolds stress models

The Reynolds stress components can be divided into an isotropic part and
an anisotropic part, which is simply defined as the deviation from the isotropic part. The Reynolds stress anisotropy is defined as

aij =

uiu j

2
ij
k
3

21

(4.1)

Formulation of an explicit algebraic Reynolds stress model

Also a transport equation for the Reynolds stress anisotropy tensor can be
construed in a similar was as the transport equation (3.1) for the Reynolds
stress component tensor. According to reference [8] it can be presented in
a rotating Cartesian coordinate system as

k Daij 1 uiu j ul uiu j kul


uiu j P
Pij ij ij

+
+ Cij( a )
=
1 +
Dt
xl
k xl
k


(4.2)
where the dissipation rate tensor ij and the redistribution tensor ij need
to be modelled. On the other hand, the production terms Pij and P = Pii 2
and the Coriolis term Cij( a ) do not need any modelling since they can be
calculated directly from the Reynolds stress tensor.
In flows where the Reynolds stress anisotropy varies slowly in time and
space, the transport equation for the Reynolds stress anisotropy tensor is
reduced to an implicit algebraic relation. Many inhomogeneous flows of engineering interest consist of a steady flow and equilibrium turbulence assumption can thus be made. Therefore the convection and diffusion of the
Reynolds stress anisotropy may be neglected. This is equivalent to the assumption made in reference [27] that the convection and diffusion of each
Reynolds stress component scale with the convection and diffusion of the
turbulent kinetic energy. This is indeed the traditional ARSM idea, to neglect convection and diffusion terms in the exact transport equation for the
Reynolds stress anisotropy. This means that the left-side terms in the
equation (4.2) can be neglected and set to zero

k Daij 1 u iu j u l u iu j ku l

=0
Dt
xl
k xl

(4.3)

The convection term Daij Dt is exactly zero for all stationary parallel mean
flows, such as fully developed channel and pipe flows. For inhomogeneous
flows the assumption of negligible diffusion effects can cause problems,
particularly in regions where the production term is small or where the inhomogeneity is strong. However, the ARSM assumption incorporates in a
natural way the effects of rotation, effects of streamline curvature and
three-dimensionality of the flow. The assumption results in an implicit alge-

22

Formulation of an explicit algebraic Reynolds stress model

braic equation for the Reynolds stress anisotropy tensor, which can be
written as

u iu j
k

P Pij ij ij
+ + Cij( a )
1 =

(4.4)

The production terms Pij and P = Pii 2 and the Coriolis term Cij( a ) do not
need any modelling since they can be calculated directly from the Reynolds
stress tensor. However, the dissipation rate tensor ij and the redistribution

ij need to be modelled. In a non-rotating coordinate system, the production term is normally written as implied in the equation (3.1). It is rewritten
here for the sake of completeness as

Pij = uiu k

u j
xk

u j u k

u i
xk

(4.5)

To illustrate the natural way in which rotational effects enter in this type of
formalism it is convenient to split the mean velocity gradient tensor into a
mean strain and a mean vorticity tensor. Symbols S ij and ij are used
here to represent these tensors, normalised with the turbulent time-scale as

u u j
S ij = i +
2 x j xi

u u j
ij = i
2 x j xi

(4.6)

where the turbulent time-scale is defined as

(4.7)

A consistent formulation of the equation (4.4) can then be obtained by replacing the mean vorticity tensor by the absolute vorticity tensor. According
to reference [8], this can formally be done introducing the absolute mean
vorticity tensor

*ij = ij + ijS

23

(4.8)

Formulation of an explicit algebraic Reynolds stress model

where the system rotation tensor ijS is defined as

ijS = jik kS

(4.9)

where ijk is the permutation tensor and kS is the constant rotation rate
vector of the system. The permutation tensor is unity when the indexes are
123, 231 or 312. It is minus unity when the indexes are 213, 321 or 132. In
other cases it is zero. Now, it is possible to rewrite the production term (4.5)
using the equations (4.1), (4.6) and (4.8). It is presented here normalised
by the dissipation rate as

4
= S ij (aik S kj + S ik akj ) + aik *kj + *ik akj

Pij

(4.10)

The Coriolis term Cij( a ) arises from the transformation of the convection
term. It is presented in this text for completeness, even though it was not
programmed into the FINFLO, as the EARSM was only implemented into a
non-rotating coordinate system. For rotational purposes, FINFLO uses a
semi-rotational formulation, which is presented in reference [28]. The
Coriolis term can be written according to reference [8] as

Cij(a ) = aik kjS ikS akj

(4.11)

For the present modelling purpose the dissipation rate tensor is assumed to
be isotropic. The equation is written here normalised by the dissipation rate

as
ij 2
= ij
3

(4.12)

The redistribution term is modelled in two subparts, slow and fast redistribution, as proposed in reference [8]. According to reference [29], the slow
redistribution rate can be though to be linear in terms of the anisotropy tensor as

ij( s )
= C1aij

24

(4.13)

Formulation of an explicit algebraic Reynolds stress model

where C1 is a model constant. For the rapid redistribution rate the general
linear Launder, Reece and Rodi (LRR) model [24] is chosen. It is normally
written for a non-rotating system as

2
C2 + 8
30C 2 2 ui u j
+
k
Pij ij P
11
3
55

x j xi
8C 2 2
2

Dij ij P
11
3

ij( r ) =

(4.14)

where the C 2 is a model constant and Dij is calculated as

Dij = uiu k

u k
u
u j u k k
x j
xi

(4.15)

A simple way of obtaining a consistent, frame-independent formulation of


the rapid pressure-strain rate model is to apply the same methodology as
for the production term, i.e. to use the mean strain and vorticity tensors defined in equation (4.4.) and to normalise by . The result can be written as

ij( r )

=
+

4
9C 2 + 6
2

S ij +
aik S kj + S ik a kj a km S mk ij
5
11
3

7C 2 10
(aik *kj *ik akj )
11

(4.16)

When the equations for the production term (4.5), the Coriolis term (4.11),
the dissipation term (4.12), the slow redistribution term (4.13) and the fast
redistribution terms (4.16) are inserted to the equation (4.4), the implicit algebraic equation for the Reynolds stress anisotropy tensor is then obtained
as

C1 1 + aij

8
S ij
15
5 9C 2
11

7C 2 + 1
(aik kjR ikR akj )
11
2

aik S kj + S ik akj ij aik S ki


3

(4.17)

Here the effective mean vorticity tensor ijR is dependent on the model
constant C 2 and therefore on the choice of the redistribution model.

25

Formulation of an explicit algebraic Reynolds stress model

The effective mean vorticity tensor ijR of the equation (4.17) can be written
as

ijR = *ij +

11
7C 2 + 12 S
ij
ijS = ij +
7C 2 + 1
7C 2 + 1

(4.18)

It should be noted, however, that in the present FINFLO formulation the


effective mean rotation rate tensor is the same as the vorticity tensor

ijR = ij , for the reasons discussed earlier. The dissipation rate must be
modelled by a transport equation similar to the one used by the twoequation eddy-viscosity models. It should be noted that equation (4.17) represents a non-linear relation, since the production to dissipation ratio is defined as

P
aik S ki

(4.19)

The algebraic Reynolds stress model has been formulated in the form of
equation (4.17). In the next section this is further simplified.

4.2

Explicit algebraic Reynolds stress model

The implicit ARSM relation for the Reynolds stress anisotropy tensor has
been found to be numerically and computationally cumbersome since there
is no diffusion or damping present in the equations. This usually means
convergence problems. In many applications the computational effort has
been found to be excessively large and the benefits of using ARSM instead
of the full Reynolds stress transport form are then successively lost, as
Wallin discusses [8]. Therefore a more simple and straightforward way to
calculate the Reynolds stress anisotropy is needed. A model, where the
Reynolds stresses are explicitly related to the mean flow field, is called an
explicit algebraic Reynolds stress model. It is much more numerically robust and has been found to have almost a negligible effect on the computational effort as compared to the linear k or the linear k model.
The procedure used here to obtain an explicit form is to avoid the nonlinearity by considering the production to dissipation ratio P as an extra

26

Formulation of an explicit algebraic Reynolds stress model

unknown. The resulting linear equation system can then be formally written
as

L ij (akl , S kl , klR , P ) = 0

(4.20)

and may, in principal, be solved directly. However, by the inspection of the


equation (4.17) it is realised that the anisotropy tensor is dependent on only
two other tensors, S ij and ijR , which can be used to form a complete base
for the anisotropy. The most general form for aij in terms of S ij and ijR
consists of ten tensorially independent groups to which all higher-order tensor combinations can be reduced with the aid of the Caley-Hamilton theorem. Various techniques have been developed with different level of simplifications in references [26], [30], [31], [32] and [33], for example. The Reynolds stress anisotropy tensor is, however, written here following reference
[8] as

1
1

a = 1S + 2 S 2 II S I + 3 2 II I + 4 (S S )
3
3

2
2

+ 5 S 2 S 2 + 6 S 2 + 2S IVI + 7 S 2 2 + 2S 2 VI (4.21)
3
3

+ 8 SS 2 S 2 S + 9 S 2 2S + 10 S 2 2 2S 2

For simplicity, boldface has been used to denote second-rank tensors as

a = aij , S = S ij and = ijR . The inner product of two matrices is defined


as (SS )ij (S 2 )ij S ik S kj and I represents the identity matrix. The
coefficients may be functions of the five independent invariants of S
and , which can be written as
II S = trace(S 2 ) = S ij S ji

II = trace( 2 ) = ij ji

III = trace(S 3 ) = S ij S jk S ki

IV = trace(S ) = S ij jk ki
2

V = trace(S 2 2 ) = S ij S jk kl li
Other scalar parameters may also be involved.

27

(4.22)

Formulation of an explicit algebraic Reynolds stress model

The value of C 2 in the rapid redistribution model, the equation (4.14), was
originally suggested to be 0.4 by Launder et. al. [24], but more recent
studies, such as [34] and [35], have suggested a higher value close to 5 9 .
This means that the last term in the ARSM equation (4.17) is quite small, if
not zero. By setting C 2 = 5 / 9 the last term of the equation (4.17) is exactly zero and a simplified but still implicit equation is obtained as

P
8
4

C1 1 + a = S + (a a )

15
9

(4.23)

The boldface notation of the tensor polynomial (4.21) has been used in
equation (4.23) again and will be used throughout the text whenever practical. The removal of the last term in the equation (4.17) gives a substantial
simplification of the solution, especially in three-dimensional mean flow.
The full model without the simplification from the equation (4.17) to the
equation (4.23) is further discussed in section 4.5.
The simplified implicit algebraic Reynolds stress equation (4.23) is then
multiplied by 9 4 to get a similar formulation as in reference [8]:

6
Na = S + (a + a )
5

(4.24)

where N is thus related to the production to dissipation ratio as

N = C1 +

9 P
4

(4.25)

where

C1 =

9
(C1 1)
4

(4.26)

The so-called Rotta coefficient, C1 , is here set to 1.8 . With the simplifications made the equation system becomes quasi-linear, as the tensor equation for a is linear and the corresponding scalar equation for N is nonlinear.

28

Formulation of an explicit algebraic Reynolds stress model

The procedure to solve this equation system is the following:


1)

The general form of the anisotropy, the tensor polynomial (4.21), is inserted into the simplified ARSM equation (4.24) where N is not yet
determined. Now N is considered to be a known parameter. This results in a linear equation system for the coefficients.

2)

The linear equation system can be solved using the fact that higherorder tensor groups can be reduced with the aid of the CayleyHamilton theorem. The solution is unequivocal because the ten groups
in the general form (4.21) form a complete basis for the system. The
solution consist of many pages of complicated tensor algebra and it is
not essential to reproduce it here. The -coefficients are now functions
of N , or the production to dissipation ratio.

3)

The next step is to formulate and solve the non-linear scalar equation
for N . This is done by inserting the solution of the coefficients into
the tensor polynomial (4.21) for the Reynolds stress anisotropy tensor.

4)

The final step is to insert the resulting equation to the simplified ARSM
equation (4.24). Then a non-linear scalar equation for N is obtained,
which for a general three-dimensional flow field is of sixth order.

The solution for the simplified ARSM equation (4.24) is presented in the
following section 4.3 for two- and 4.4 for three-dimensional mean flows. In
section 4.5, the solution for the model without the simplification made from
the equation (4.17) to (4.23) is also presented.

4.3

Simplified EARSM for two-dimensional mean flow

For two-dimensional mean flows the simplified solution of the ARSM equation is reduced to only two non-zero coefficients, which can be expressed
according to reference [8] as

1 =

N
6
2
5 N 2 II

6
1
4 =
2
5 N 2 II

29

(4.27)

Formulation of an explicit algebraic Reynolds stress model

In the equation (4.27) it is clearly seen that the denominator cannot become
singular since II is always negative. The non-linear equation for N in twodimensional mean flow can be derived by inserting the tensor polynomial
(4.21) for the Reynolds stress anisotropy tensor with the coefficients from
the equation (4.27), into the definition of N, equation (4.24). The resulting
equation is cubic and can be written following reference [8] as

27

N 3 C1N 2
II S + 2 II N + 2C1II = 0
10

(4.28)

The equation can be solved in a closed form with the solution for the positive root being

1
1
C1
3
3
P
P
P
P
sign P1 P2
+
+
1 +
2
1
2
3

N =
1

C1
P1

+ 2(P12 P2 ) cos arccos


2

3
P
P

3
2
1

, P2 0
(4.29)

, P2 < 0

where the arccos-function should return an angle between 0 and . The

P1 and P2 can be calculated as

C12

9
2
P1 =
+
II S II C1
3
27 20

C1

9
2
P2 = P12
+
II S + II
10
3
9

(4.30)

It can easily be shown that N remains real and positive for all possible
values of II S and II . When the N is known from the equation (4.29), the
production to dissipation ratio is calculated from the equation (4.25) and the
system is completely solved. The formulation of this simplified twodimensional EARSM has been included in this text for the sake of completeness. The simplified two-dimensional model, and therefore this section, will also be used in the formulation of the simplified three-dimensional
EARSM in a manner to be presented in the next section. However, the
simplified two-dimensional EARSM is not discussed further.

30

Formulation of an explicit algebraic Reynolds stress model

4.4

Simplified EARSM for three-dimensional mean


flow

According to reference [8], for general three-dimensional mean flows the


simplified solution of the ARSM for the coefficients is written as

1 =

N 2 N 2 7 II
Q

3 =

12 N 1 IV
Q

2 N 2 2 II
Q
6N
6 =
Q
6
9 =
Q

4 =

(4.31)

where all the other -coefficients are identically zero. The denominator Q
is calculated as

Q=

5 2
(N 2 II ) (2 N 2 II )
6

(4.32)

The denominator can not become singular since II is always negative.


The non-linear equation for N and for P (4.25) is obtained by introducing the solution for the -coefficients (4.31) for the Reynolds stress anisotropy tensor polynomial (4.21) and then into the definition of the simplified
ARSM equation (4.24) for N . The resulting equation is of sixth order and
can be written, according to reference [8], as

5
5
27

N 6 C1N 5
II S + II N 4 + C1II N 3
2
2
10

189
81
81 2

+ II 2 +
II S II V N 2 C1II 2 N
IV = 0
20
5
5

(4.33)

Equation (4.33) cannot be solved in a closed form, which means that this
approach gets into an inextricable dilemma and a recourse is necessary. A

31

Formulation of an explicit algebraic Reynolds stress model

possible way to continue is to get the non-linear equation (4.28) for N in


two-dimensional mean flow and use it as a first approximation. In twodimensional mean-flows there are only two independent invariants, II S and

II . The other invariants are III S = IV = 0 and V = II S II 2 . The resulting equation (4.28) is rewritten here for completeness

27

N c3 C1N c2
II S + 2 II N c + 2C1II = 0
10

(4.34)

The symbol N c is used instead of N to emphasise the fact that equation


(4.34) is an equation for two-dimensional mean flows. The solution for the
positive root is already given by the equations (4.29) and (4.30). According
to reference [8], a further improvement of the approximation of N by the
two-dimensional mean flow solution is also possible by making a perturbation solution of the three-dimensional equation. This can be done by perturbing the IV and V invariants around the two-dimensional solution as

IV = 1
V=

(4.35)

1
II S II + 2
2

and assuming that 1 and 2 are independent of each other. When the
equation (4.35) is inserted into the three-dimensional equation (4.33) for N
the result is after some rearrangement

) (

162 1 + 2 N c2
N = Nc +
+ O 12 , 22 , 1 2
D

(4.36)

where the denominator is given by

D = 20 N c4 N c C1 II (10 N c3 + 15C1N c2 ) + 10C1II 2


2

(4.37)

It is obvious that the denominator D will always remain positive, because

N c C1 and II 0 . The production to dissipation ratio is found from the


equation (4.25) and the system is completely solved.

32

Formulation of an explicit algebraic Reynolds stress model

It could also be considered that N was kept implicit during the iteration
procedure to obtain a steady-state solution. This would perhaps give
slightly more accurate results according to Wallin [8]. The consequences of
this on the stability and numerical behaviour are not, however, known and
therefore this approach should be avoided (Ref. [8]).

4.5

Solution of the general quasi-linear ARSM equation

The removal of the last term in the equation (4.17) lead to a substantial
simplification of the solution, especially in three-dimensional mean flow.
The simplification was done by setting C 2 = 5 / 9 in the implicit equation
(4.17). The simplified implicit equation (4.23) was obtained, which is rewritten here as a reminder

P
8
4

C1 1 + a = S + (a a )

15
9

(4.38)

In this section the full non-simplified model is formulated. The implicit


ARSM equation (4.17) is written here in a more compact form without the
simplification as

Na = A1S + (a a ) A2 aS + Sa trace(aS )I
3

(4.39)

where N is related to the production to dissipation ratio as

N = A3 + A4

(4.40)

The A coefficients are calculated as

11(C1 1)
7C 2 + 1
11
A4 =
7C 2 + 1

88
11(7C 2 + 1)
5 9C 2
A2 =
7C 2 + 1
A1 =

A3 =

33

(4.41)

Formulation of an explicit algebraic Reynolds stress model

The values of the A-coefficients can be found from the table 4.1 for the different models.
Table 4.1. The coefficients of the general ARSM for the different models.

A1

A2

A3

A4

Current model (C1 = 1.8, C2 = 5/9)

1.20

1.80

2.25

Original LRR (C1 = 1.5, C2 = 0.4) [24]

1.54

0.37

1.45

2.89

Linearized SSG [25]

1.22

0.47

0.88

2.37

Gatski & Speziale without regularisation [33]

1.22

0.47

5.36

Model

According to reference [8], the solution of the linear equation system where

N is assumed to be a known variable can be formulated as

N = A1 1 + J A2 H

(4.42)

which can also be written in the standard form for linear equation systems
as

(N

J + A2 H ) = A1 1

(4.43)

According to reference [8], the matrixes H and J are written for a threedimensional mean flow as
0

2
0

0
0
H =
0
0

0
0

1
3

II S
0
0

2
3

2
3

II
0

0
0

0
0

IV
2 II

II S

1
2

II S

13 V
IV
1
3

0
0

0
0

III S

IV

II

II S

III S
0

1
2

1
3

0
0

0
0

0
0

1
0

0
0

0
0

0
0

0
0

34

1
1
3 V 6 II S II (4.44)

23 IV

II
3

1
3 II S

0
0
0

Formulation of an explicit algebraic Reynolds stress model

and
0

0
0

1
0
J =
0
0

0
0

0 0 II
0 0
0
0 0
0

0
II
2 II S

0
0
0
1
2

0
0
0

2V II S II
2 IV
0

II 2
0
2 IV

0
II

0
0

0
0

0 0
1 0

0
0

0
0

0 0
0 0

3
0

0
3

0
0

0
0

II S
0

2 II
0

0 0
0 0

0
0

0
0

0
1

0
0

0
0

0
0

0 0

II
0

1
2

2
II

2 II S II 2V

0
(4.45)

2 II

The solution for a three-dimensional mean flow can then be obtained. It is


written here for the coefficients after rearrangement as

A1 N
30 A2 IV 21NII 2 A23 III S + 6 N 3 3 A22 II S N
2Q
AA
2 = 1 2 6 A2 IV + 12 NII + 2 A23 III S 6 N 3 + 3 A22 II S N
Q
3A
3 = 1 2 A22 III S + 3NA2 II S + 6 IV
Q
A
4 = 1 2 A23 III S + 3 A22 II S N + 6 A2 IV 6 NII + 3N 3
Q

1 =

5 =

9 A1 A2 N 2
Q

9 A1 N 2
Q
18 A1 A2 N
7 =
Q

6 =

9 A1 A22 N
Q
9A N
9 = 1
Q
10 = 0

8 =

(4.46)

35

Formulation of an explicit algebraic Reynolds stress model

The denominator Q in (4.46) is calculated as

15
Q = 3 N 5 + II A22 II S N 3 + (21A2 IV A23 III S )N 2
2

2
2
(4.47)
+ (3II 2 8 II S II A22 + 24 A22V + A24 II S2 )N + A25 II S III S
3
+ 2 A23 II S IV 2 A23 II II S 6 A2 II IV
The equation for N can also be derived for three-dimensional mean flow
and is a sixth order polynomial equation. According to reference [8] the expression is, however, much more complicated than the equation (4.33) and
of small practical interest. As for the solution of the simplified ARSM equation, the two-dimensional solution of N can also be used here as a first
approximation, as Wallin discusses [8]. The exact formulation of the solution of N is, however, clearly beyond the scope of this text.
It is noted that the general quasi-linear two-dimensional model has been
implemented into the two-dimensional version of the FINFLO. The test calculations with the model have been made using the SSG-coefficients of
reference [25]. There were no significant improvements of the results with
the general quasi-linear model compared to the results of the simplified
two-dimensional EARSM, as can be seen from references [36] and [37].
This is not necessarily true, though, for the three-dimensional model.

36

Chapter 5
Programming approach and written subroutines

In this chapter the programming of the EARSM is discussed. The turbulence model was implemented using Fortran 77 and 90 into the NavierStokes solver called FINFLO for three-dimensional flows. The implemented
version of the turbulence model is the simplified EARSM for threedimensional flows, which is described in section 4.4. In the beginning of the
chapter the five subroutines that were written are presented and discussed.
A brief description of the modifications of the boundary routines is also presented.

5.1

Subroutine EARSM1

The main core of the EARSM is a subroutine called EARSM1 where the
effective eddy-viscosity coefficient and the extra anisotropy components
are calculated. The source code is presented in the appendix A. A brief description of the main elements of the subroutine is presented here.
The EARSM was implemented to use the k model as a background.
The contribution of the EARSM is most easily implemented when the Reynolds stress components are divided into two parts. The isotropic part of

37

Programming approach and written subroutines

the Reynolds stresses is handled as an eddy-viscosity by the k model.


The model also calculates the ratio of the dissipation rate to the turbulent
kinetic energy, . The dissipation rate , which is needed in the solution
of the EARSM, is easy to calculate as = * k , where * = 0.09 . The
isotropic part of the Reynolds stresses can be written in terms of an effective eddy-viscosity coefficient, which is defined as

C eff =

1
(1 + II 6 )
2

(5.1)

Note that also a part of the 6 term is calculated with the 1 term. By this
formulation the effective eddy-viscosity coefficient adapts to the local strainand vorticity field by the invariants. This is the fundamental improvement of
the EARSM, because in the linear eddy-viscosity models the eddy-viscosity
coefficient can be though of being constant. This improvement removes the
need for the SST-limitation of the k model. The kinematic eddyviscosity used by the k model is calculated from the eddy-viscosity coefficient as

T = C eff k

(5.2)

The second, anisotropic, part of the Reynolds stress components is defined


as the deviation from the isotropic part. Using this formulation the Reynolds
stress components can be expressed as a sum of the isotropic and the anisotropic part as

u iu j = k ij 2C eff S ij + aij(ex )
3

(5.3)

The extra anisotropy tensor is written, using the boldface to represent the
second-rank tensors, as

a (ex )

= 3 S 2 II S I + 4 (S S )
3

+ 6 S + S II S IVI + 9 (S 2 2S )
3

38

(5.4)

Programming approach and written subroutines

where the coefficients are calculated from the equation (4.31). This way
the extra anisotropy can be added to the equations and to the calculation
routines as fully explicit additional terms. This is formally done in the different subroutine called ANISFLUX, which is presented in the section 5.3.
The strain and vorticity tensors are calculated from the main velocity gradients and scaled with the turbulent time-scale as

ui u j
+
S ij =
2 x j xi

ui u j
ij =

2 x j xi

(5.5)

The turbulent time-scale is calculated in a different subroutine called


TTIMES, which is presented in detail in the section 5.2. Because the strain
tensor is symmetric and the vorticity tensor anti-symmetric, there are six
different strain and three vorticity components in a three dimensional mean
flow as

S11
S = S12
S13

S12
S 22
S 23

S13
S 23
S 33

0
= 12
13

12
0
23

13
23
0

(5.6)

The invariants of the strain and vorticity tensors are needed for the solution
of the -coefficients and for the solution of the corresponding tensor components. They are calculated as implied by the equation (4.22). The square
invariant of the strain tensor is calculated as

II S = S112 + S 222 + S 332 + 2(S122 + S132 + S 232 )

(5.7)

It can be seen that the invariant II S can not become negative. The square
invariant of the vorticity tensor is calculated as
2
2
II = 2(12
+ 13
+ 223 )

(5.8)

It is noted that the invariant II is always equal to or less than zero. The
cube invariant of the strain tensor is presented here for the sake of completeness even though it is not used by the simplified version of the
EARSM.

39

Programming approach and written subroutines

The cube invariant of the strain tensor could be calculated as

III

2
= S11 (S112 + S122 + S132 ) + S 22 (S122 + S 22
+ S 232 )

+ S 33 (S132 + S 232 + S 332 ) + 2 S12 (S11 S12 + S12 S 22 + S13 S 23 )

+ 2 S13 (S11 S13 + S12 S 23 + S13 S 33 ) + 2 S 23 (S12 S13 + S 22 S 23 + S 23 S 33 )

(5.9)

The cube invariant, where the strain tensor is linear and the vorticity tensor
is squared, is calculated as

IV

2
2
= S11 (12
+ 13
) S 22 (122 + 223 ) S33 (132 + 223 )
+ 2( S12 13 23 + S13 12 23 S 2312 13 )

(5.10)

The fourth order invariant where both the strain and the vorticity invariants
are squared is calculated as

2
2
= (S112 + S122 + S132 ) (12
+ 13
)
2
) (122 + 223 )
(S122 + S 222 + S 23

2
(S132 + S 232 + S 332 ) (13
+ 223 )
2(S12 S13 + S 22 S 23 + S 23 S 33 ) 12 13
+ 2(S11 S13 + S12 S 23 + S13 S 33 ) 12 23
2(S11 S12 + S12 S 22 + S13 S 23 ) 13 23

(5.11)

The tensor component terms of the tensor component equation (5.4) for the
anisotropy tensor to be used with 3 are presented here as an example.
They are calculated as

1
2 1

2
2
II I = 12 13 II
3
3

11
2 1

II I = 13 23
3

12
2 1

II I = 12 23
3

13
1
2 1

2
2
II I = 12 23 II
3
3

22
2 1

II I = 12 13
3

23

40

(5.12)

Programming approach and written subroutines

The sixth term (33) of the tensor needs not to be calculated because the extra anisotropy is traceless. It can thus be calculated simply as

a33 = a11 a 22

(5.13)

The tensor component terms to be used with 4 , 6 and 9 are not presented here because of their complexity. Their formulation can be read
from appendix A where the subroutine EARSM1 is presented. It is thus unnecessary to repeat the whole calculation routine of the chapter four. The
reader is encouraged to consult the appendix A for further details.
It is noted that in the FINFLO formalism a different symbol is used for the
extra anisotropy tensor. The extra anisotropy tensor is referenced in the
code as b(ex), which is simply one half of the a(ex) used here in this text. The
relationship aij = 2bij is confusing, but it is based on the historical aspects
of the FINFLO notation and was thus used in the programming.
The correct programming of the subroutine was tested several times using
different approaches.

5.2

Subroutine TTIMES

The turbulent time-scale is calculated in a different subroutine called


TTIMES, which is presented in the appendix B. The current version of the
EARSM has only been tested with the k model but also a k formulation of the turbulent time-scale is programmed in the subroutine. However, the details of the k formulation are not presented here. The actual
turbulent time-scale is calculated with the k model as

tur =

1
*

(5.14)

where the model coefficient * = 0.09 . The viscous time-scale, which is


also called the Kolmogorov scale, is calculated as

vis = C

*k

41

(5.15)

Programming approach and written subroutines

where the model coefficient C = 6.0 . The actual turbulent time-scale, defined by the equation (5.15), goes to zero as the wall is approached because the turbulent kinetic energy vanishes. It would be unacceptable to let
the numerical turbulent time-scale also go to zero, because it would zero
the strain and vorticity tensors that are scaled with the time-scale. The
Kolmogorov scale stays positive but small when the wall is approached.
Therefore the lower limit of the turbulent time-scale is given by the Kolmogorov scale as

= max( tur , vis )

(5.16)

This is, in fact, the only near-wall correction used with the present form of
the EARSM.

5.3

Subroutine ANISFLUX

The FINFLO calculates the fluxes using the isotropic part of the Reynolds
stresses with the k model running on the background. However, also
the anisotropic parts of the Reynolds stress components have to be added
to the fluxes (see the equation 5.3). This is calculated in a subroutine called
ANISFLUX, which is presented completely in the appendix C.
First, the cell boundary values of the velocity, turbulent kinetic energy and
extra anisotropy components are calculated as simple arithmetic means of
the values on both sides of the boundary. Then the anisotropic parts of the
momentum fluxes Fi are calculated as

Fi = k Sn j aij

(5.17)

where k is the mean value of turbulent kinetic energy, S is the cell face
area, n j is the cell face unit normal vector component and aij is the mean
extra anisotropy component. The anisotropic part of the energy flux Fe is
calculated as

Fe = k Sni u j aij

42

(5.18)

Programming approach and written subroutines

where u j is the cell boundary value of the corresponding velocity component. At the end of the subprogram the anisotropic parts of the fluxes are
added to the momentum and energy fluxes. The reader is encouraged to
consult the appendix C for further details. This subprogram was also tested
separately for correct programming.

5.4

Subroutine FLUXP

During the programming and testing of the EARSM it also became apparent that a subroutine called FLUXP was to be modified. The subroutine calculates the circumferential fluxes for an axisymmetric flow case, and a correction of the fluxes similar to the subroutine ANISFLUX was also needed
in the subroutine FLUXP. As the discussion of the previous section also
applies to this section, no further details of the subroutine are presented.
The reader is encouraged to consult the appendix D for further details.

5.5

Subroutine PEEKOO

The production of turbulence is calculated separately with the EARSM as


has been discussed in the previous chapters. It is calculated in a subroutine
called PEEKOO, which is presented in the appendix E.
The Reynolds stress tensor components are calculated as

u u
2 u k
2

uiu j = T i + j
ij + k ij + aij(ex )
3

x j xi 3 xk

(5.19)

The production of turbulence is calculated as

P = u iu j

u i
u j

(5.20)

In the subroutine also a possibility to force the flow laminar has been included as can been noted from the appendix E.

43

Programming approach and written subroutines

5.6

Subroutine VELGRAD

The values of the mean flow velocity gradients, the strain invariant and vorticity components are calculated and stored in a subroutine called
VELGRAD. It is presented in the appendix F. The gradient of the velocity
component u m in the direction of xn is basically calculated for a cell

( i, j, k )

as

u m
xn

(u m S n )i +1 2 (u m S n )i 1 2
Vi , j , k
(u m S n ) j +1 2 (u m S n ) j 1 2
+

(u m S n )k +1 2 (u m S n )k 1 2 ]

(5.21)

where the Vi , j , k is the volume of the cell, the (u m )i + 1 2 is the velocity com-

ponent on the cell wall and the (S n )i + 1 2 is the cell face area perpendicular

to the xn . The velocity component values of the cell walls are calculated as
a simple arithmetic mean of the centre values of the adjacent cells.
The calculation procedure is based on the finite volume method and depends on the grid geometry formulation, and further details are thus beyond
the scope of this text. If the reader is interested in the calculation method,
he or she is encouraged to get acquainted with the appendix F.

5.7

Boundary routines

There are ten different types of boundary conditions in the FINFLO [11]:
-

CON (connectivity)

PER (periodic)

MIR (mirror)

INL (inlet)

OUT (outlet)

EXT (external)

SOL (solid)

MOV (moving solid)

ROT (rotating solid)

SNG (singularity)

44

Programming approach and written subroutines

Among the other quantities, also the extra anisotropy has to be connected
to another block face (routines CON and PER). Also, the extra anisotropy
has to be extrapolated to the ghost cells to ensure that the momentum and
energy fluxes are not miscalculated at the edges of each block (routines
MIR, INL, OUT and EXT). There are two columns or rows of ghost cells
around each block as shown in the figure 5.1.

Figure 5.1. Schematic picture of the ghost cells.

45

Chapter 6
Test cases

In this chapter the test cases are presented and discussed. Four test cases
were selected to verify that the EARSM has been correctly implemented.
The test cases were calculated with Roes flux splitting using a third order
upwind biased MUSCL extrapolation.
First, the effective eddy-viscosity coefficient C eff and its connection with
the k BSL model were tested. A two-dimensional boundary layer over
a flat plate was chosen, as it is an easy starting point.
Secondly, the second order effects of the EARSM were tested with a fully
developed flow inside a rectangular duct. Secondary flows occur in turbulent flows along any noncircular duct due to the anisotropy of the Reynolds
stresses. The test case was thus used to verify the implementation of the
second order terms, namely 3 (S 2 II S I 3) and 4 (S S ) .

Thirdly, the third order effects of the implemented EARSM were tested with
a fully developed flow inside a rotating pipe. This is indeed a very good test
case, because the three-dimensional effects of the flow field are purely turbulence driven. This behaviour is very extreme since in most cases the
three-dimensional effects driven by turbulence are quite weak compared to

47

Test cases

the three-dimensional effects driven by mean momentum forces. This test


case was used to verify the correct implementation of the third order term,
namely 6 (S + S II S 2 IVI 3) .

The fourth test case selected to test the model was a flow over a cylinder
part of which rotates. This test case was selected to further verify the sound
implementation of the third order terms, which were already tested with a
fully developed flow. However, this fourth case is fundamentally different
than the third case because the flow is developing due to the sudden strain
caused by the moving surface of the cylinder. Also the transient behaviour
of the model is thus also tested.

6.1

Flow over a flat plate

A boundary layer flow above a flat plate with a zero pressure gradient was
chosen to be the first test case for the model. This is a very well-known flow
case with a variety of available approximate solutions, which makes it an
easy starting point. A large number of experimental measurement results
also exist, of which maybe the best-known are the ones of P. S. Klebanoff
presented in reference [38]. The results are presented in a point where the
local Reynolds number from the front edge of the plate is 4.2 million.

6.1.1 Grid
The grid consisted of two blocks. The flat plate itself was located on the
bottom of the second block. The first block was used to provide a correct
steady inflow in front of the plate. It consisted of 8 x 96 x 1 cells in the direction of x, y and z, respectively. The second block consisted of 128 x 96 x
1 cells. The grid is presented in figure 6.1.

Figure 6.1. The grid of the flat plate model. The first block is red and the
second green.

48

Test cases

The length, height and width of the first block were unity. The size of the
smallest cell on the lower right corner of the block, just before the beginning
of the plate was 0,01605 x 0,00001 x 1 units. The grid was geometrically
expanded to both directions using a constant ratio between two adjacent
cells.
The height and width of the second block were also unity. The length of the
block was five units. The second block used the same size of the smallest
cell just above the beginning of the plate on the lower left corner of the
block and the grid was geometrically expanded to both directions. At the
point where the local Reynolds number was 4.2 million the local xcoordinate measured from the beginning of the plate was 1.922, which was
the 75th cell of the block in x-direction. The results are drawn at this point.
The Reynolds number at the end of the plate was 10.927 million.
The non-dimensional y+ value of the lowest cell is defined as

y+

u y

(6.1)

(6.2)

The friction velocity is by definition as

And the wall shear stress is defined as

1
w U 2 c f
2

(6.3)

The y+ can thus be calculated as

y + = ReL

cf y
2 L

(6.4)

The non-dimensional y-coordinate is thus y+ = 0.8. According to reference


[39], the grid must get some points inside the sublayer, which is usually
considered to reach from the wall to a y+ value of five. It is thus acknowl-

49

Test cases

edged that values of the y+ around unity are sufficiently small. The first cells
of the grid inside the boundary layer were thus small enough to capture the
large gradients of quantities that appear in the sublayer.

6.1.2 Boundary conditions


The numbering of the block faces is done in the FINFLO formulation as
shown in figure 6.2.

Figure 6.2. The numbering of the block faces used in FINFLO.


The boundary conditions that were used are presented in table 6.1.
Table 6.1. The boundary conditions for the flat plate model.
Face

Block 1

EXT

MIR

MIR

CON

EXT

MIR

Block 2

CON

SOL

MIR

EXT

EXT

MIR

Where EXT means external or free stream condition, MIR symmetric mirror
image flow, SOL solid wall and CON connection between the face number
four of the first block and the face number one of the second block.

50

Test cases

6.1.3 Results
The calculation was carried out using Courant numbers of six on the fine
grid level and nine on the coarser grid levels. Three multi-grid levels were
used and it took about 8 000 iteration cycles for the convergence to be
reached. The criterion for the convergence was that the residuals of each
component of the momentum vector and the turbulence quantities did not
decrease any more in time. The total pressure, mass, kinetic energy and
turbulent kinetic energy remained also constant when the convergence was
reached. The same criteria was also used in the other test cases.
First, velocity profiles are shown in figure 6.3. The experiment results are
taken from the Klebanoffs report, reference [38]. The velocity profiles with
both the k BSL and the EARSM follow the experimental results closely
deviating slightly from each other only at the edge of the boundary layer,
because the k BSL has not really been fine-tuned to work perfectly
with the EARSM. The present model gives slightly wrong results in the areas where the strain dissappears, i.e. II S 0 , because the diffusion term
of the anisotropy tensor has been neglected. The fine-tuning of the k
BSL model was beyond the scope of the thesis. It has, however, been
started at the moment of writing.
1.0
0.9
0.8

u/U0

0.7
0.6
0.5
0.4
0.3
BSL

0.2

EARSM

0.1

Experiment

0.0
0.0

0.1

0.2

0.3

0.4

0.5
y/

0.6

0.7

0.8

0.9

1.0

Figure 6.3. Velocity profiles with the EARSM and the k BSL compared
to the experimental of reference [38]. Rex = 4.2 10 6 .

51

Test cases

Production of turbulence is presented in figure 6.4. Close to the wall the


EARSM produces slightly less turbulence than the k BSL.
2.0E+05
1.8E+05
1.6E+05

BSL

1.4E+05

EARSM

P [m2/s3]

1.2E+05
1.0E+05
8.0E+04
6.0E+04
4.0E+04
2.0E+04
0.0E+00
0.0000

0.0005

0.0010

0.0015

0.0020

0.0025

0.0030

0.0035

0.0040

0.0045

0.0050

y [m]

Figure 6.4. Production of turbulence with the EARSM and the k BSL.
Rex = 4.2 10 6 .
The turbulent kinetic energy is presented in figure 6.5. The EARSM predicts
a slightly narrower boundary layer than the k BSL. The same conclusion can, of course, be made from figure 6.3.
25

BSL

20
EARSM

k [m /s ]

15

10

0
0.000

0.005

0.010

0.015

0.020

0.025

0.030

0.035

y [m]

Figure 6.5. Turbulent kinetic energy with the EARSM and the k BSL.
Rex = 4.2 10 6 .

52

Test cases

The Reynolds stress components are presented in figures 6.6 and 6.7 calculated with the EARSM. The agreement with the Klebanoffs experimental
results [38] is very good the only thing not predicted correctly being the
peak of u u very close to the wall. With the present EARSM, no wall corrections or damping functions are used to correct this minor shortcoming.
Such corrections have been tested with the two-dimensional version of the
FINFLO, but the effect on the flow field in general is only minor.
0.10
u/U, Experiment

0.09

v/U, Experiment
0.08

w/U, Experiment
u/U, EARSM

0.07
u/U, v/U, w/U

v/U, EARSM
0.06

w/U, EARSM

0.05
0.04
0.03
0.02
0.01
0.00
0.0

0.2

0.4

0.6

y/

0.8

1.0

1.2

1.4

Figure 6.6. The distributions of the turbulent intensities with the EARSM
compared to the experimental results of reference [38]. Rex = 4.2 10 6 .
0.0030

0.0025

0.0020
2

2u 0.0015
v/
U
Experiment
0.0010
EARSM
0.0005

0.0000
0.0

0.2

0.4

0.6

0.8

1.0

1.2

y/

Figure 6.7. The distribution of the turbulent shearing stress with the
EARSM compared to the experimental results of reference [38].
Rex = 4.2 10 6 .

53

Test cases

The skin friction coefficients of the plate are presented in figure 6.8. The
experimental results are from reference [40]. The k BSL gives almost
identical results with the experiments. The curve called White is calculated
with an equation for a turbulent boundary layer from reference [41]. The
transition point is fixed at the front edge of the plate. The equation can be
written as

Cf =

0.455
[ln(0.06 Rex )]2

(6.5)

where Rex is the local Reynolds number. The equation (6.5) gives slightly
higher skin friction coefficients than measured or calculated with k
BSL. The EARSM has not been fine-tuned yet and gives slightly smaller
skin friction coefficients.
0.0040
0.0035
0.0030

Cf

0.0025
0.0020
BSL

0.0015

EARSM
0.0010
Experiment
0.0005

White

0.0000
0.0E+00 1.0E+06 2.0E+06 3.0E+06 4.0E+06 5.0E+06 6.0E+06 7.0E+06 8.0E+06 9.0E+06 1.0E+07

ReX

Figure 6.8. Skin friction coefficients with the EARSM and the k BSL.
The experimental results are from reference [40] and the curve White is
calculated with the equation (6.5).
The connection between the eddy-viscosity coefficient of the EARSM and
the k BSL seems to be correctly implemented on the basis of this test
case. However, the underlying k model needs some fine-tuning.

54

Test cases

6.2

Flow inside a rectangular duct

Noncircular ducts are used often in industry. It is well known that secondary
flows of Prandtls second kind occur in turbulent flows along these ducts.
The occurrence of this secondary flow pattern is of great practical interest
because it causes additional pressure losses and enhances mixing. According to reference [42], this secondary motion is generated by the imbalance between gradients of the secondary Reynolds shear stress and the
normal stress anisotropy. Although the strength of the secondary flow generated solely by the turbulence field is usually at least an order of magnitude weaker than the mean flow field, it still has profound effects on the
lower-order flow statistics.
A classical example of the secondary flow field is the fully developed turbulent flow in a straight square duct, which has been used as a test case
for the assessment of turbulence models for more than 25 years, for example [42] and [43]. The test case is considered to be in a fully developed
state and the stress transport effect and the convection term is thus of no
importance. Therefore, according to reference [43], the EARSM is a fairly
accurate approximation of the second-moment closure.
The results of the calculations are compared to the well-known experimental measurements conducted by Yokosawa, Fujita, Hirota and Iwata in reference [44].

6.2.1 Grid
Only a quarter of the duct was modelled because of symmetry. The grid
consisted of two blocks. The first block was a so-called inlet block, which
was used to give sound values for the density, momentum components,
total temperature, turbulence level and turbulent viscosity. The inlet block
consisted of 2 x 96 x 96 cells in the direction of x, y and z, respectively. The
second block, in which the actual solution was calculated, consisted of 96 x
96 x 96 cells. The grid is presented in figure 6.9.

55

Test cases

Figure 6.9. The grid of the rectangular duct. The inlet block is green and
the calculation block red. The figure has been stretched in the y-direction
for clarity.
The height and width of the inlet block were unity. Its length was four units.
The dimensions of the first cell next to the wall were 2 x 0.0003 x 0.0003
units. The height and width of the second block were also unity and its
length was 200 units. The second block used the same size of the first cell
just at the beginning of the duct. The Reynolds number referred to the axial
bulk velocity and to the width of the whole duct was 65 000. The nondimensional y+ value of the lowest cell can be calculated in this case as

cf y
2 H

y + = ReH

(6.6)

The non-dimensional y-coordinate is therefore y+ = 0.4. Therefore the first


cells of the grid inside the boundary layer were small enough.

6.2.2 Boundary conditions


The boundary conditions that were used are shown in table 6.2.
Table 6.2. Boundary conditions for the rectangular duct.
Face
Inlet block
Calculation block

INL

SOL

SOL

CON

MIR

MIR

CON

SOL

SOL

EXT

MIR

MIR

56

Test cases

Where INL mean user specified inlet condition, where density, momentum
components, total temperature, turbulence level and turbulent viscosity are
given. CON means connection between the face numbers four of the inlet
and one of the actual calculation block.

6.2.3 Results
The results of the calculations were compared to the experimental results
found from reference [44]. The Reynolds number based on the axial mean
velocity and the channel width was 65 000. The results are drawn at a station close to the outlet where the flow was fully developed.
First the axial velocity distributions in three different cross sections are presented in figures 6.10, 6.11 and 6.12. The reference value h denotes the
height or the width of the quarter of the channel.
1.0
0.9
0.8
0.7

U1/UC

0.6
0.5
0.4
0.3

Experiment

0.2

BSL
EARSM

0.1
0.0
0.0

0.1

0.2

0.3

0.4

0.5
y/h

0.6

0.7

0.8

0.9

1.0

Figure 6.10. Axial velocity distribution in section where z/h = 0.2 with the
EARSM and the k BSL compared to the experimental results of reference [44].

57

Test cases

1.0
0.9
0.8
0.7

U1/UC

0.6
0.5
0.4
0.3

Experiment

0.2

BSL
EARSM

0.1
0.0
0.0

0.1

0.2

0.3

0.4

0.5
y/h

0.6

0.7

0.8

0.9

1.0

Figure 6.11. Axial velocity distribution in section where z/h = 0.6 with the
EARSM and the k BSL compared to the experimental results of reference [44].
1.0
0.9
0.8
0.7

U1/UC

0.6
0.5
0.4
0.3

Experiment

0.2

BSL

0.1

EARSM

0.0
0.0

0.1

0.2

0.3

0.4

0.5
y/h

0.6

0.7

0.8

0.9

1.0

Figure 6.12. Axial velocity distribution in section where z/h = 1.0 with the
EARSM and the k BSL compared to the experimental results of reference [44].

58

Test cases

The velocity distributions of the secondary flow in y-direction in three different cross sections are presented in figures 6.13, 6.14 and 6.15.
0.015

0.010

U2/UC

0.005

0.000

Experiment
-0.005
BSL
EARSM
-0.010
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

y/h

Figure 6.13. Velocity distribution of the secondary flow in y-direction in


section where z/h = 0.2 with the EARSM and the k BSL compared to
the experimental results of reference [44].
0.015
Experiment
BSL

0.010

EARSM

U2/UC

0.005

0.000

-0.005

-0.010
0.0

0.1

0.2

0.3

0.4

0.5
y/h

0.6

0.7

0.8

0.9

1.0

Figure 6.14. Velocity distribution of the secondary flow in y-direction in


section where z/h = 0.6 with the EARSM and the k BSL compared to
the experimental results of reference [44].

59

Test cases

0.015
Experiment
BSL

0.010

EARSM

U2/UC

0.005

0.000

-0.005

-0.010
0.0

0.1

0.2

0.3

0.4

0.5
y/h

0.6

0.7

0.8

0.9

1.0

Figure 6.15. Velocity distribution of the secondary flow in y-direction in


section where z/h = 1.0 with the EARSM and the k BSL compared to
the experimental results of reference [44].
Thirdly the velocity distributions of the secondary flow in z-direction in two
different cross sections are presented in figures 6.16 and 6.17.
0.015
Experiment
BSL

0.010

EARSM

U3/UC

0.005

0.000

-0.005

-0.010
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

y/h

Figure 6.16. Velocity distribution of the secondary flow in z-direction in


section where z/h = 0.2 with the EARSM and the k BSL compared to
the experimental results of reference [44].

60

Test cases

0.015
Experiment
BSL

0.010

EARSM

U3/UC

0.005

0.000

-0.005

-0.010
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

y/h

Figure 6.17. Velocity distribution of the secondary flow in z-direction in


section where z/h = 0.6 with the EARSM and the k BSL compared to
the experimental results of reference [44].
The velocity distribution of the secondary flow in z-direction in not presented in the section where z/h = 1.0 because it is identically zero due to
the symmetry.
The experimental results of the secondary flow magnitudes found from reference [44] contain quite a lot of random noise. Perhaps the sample time
has not been long enough. However inaccurate the experimental results
may be, it can be concluded that the EARSM is able to predict the secondary motion right qualitatively. Reference [42] also includes results of the
numerical calculations made with a second order closure implemented in a
finite-volume/finite-difference code called TEAM-RSM. The figures are not
reproduced here in this text though because they are similar to the previous
figures and the author feels that presenting them would be futile. When
these figures are compared to the results of the EARSM, it can be noted
that the EARSM manages to produce the motion in the same manner as
the differential Reynolds stress model of reference [42]. Also the second
order terms of the EARSM thus seem to be correctly implemented.
It was also tested whether the solution is grid independent. The previous
results have been presented on the first grid level, i.e. the finest. The second grid level means that only every other grid line is considered, which

61

Test cases

means that only one eighth of the grid points are calculated. The calculation
is thus roughly eight times faster than on the first grid level. The third grid
level means that only one 64th of the grid points are calculated, the calculation being roughly 64 times faster than on the first level. There are certainly
other techniques for the testing of numerical uncertainty, but, according to
reference [39], systematic grid convergence studies are the most common,
most straightforward and arguably the most reliable.
Figure 6.18 represents the axial velocity distribution for the different grid
levels using the EARSM. The results are from the section where z/h = 0.2.
1.0
0.9
0.8
0.7

U1/UC

0.6
0.5
0.4
0.3

EARSM, Grid level 1

0.2

EARSM, Grid level 2

0.1

EARSM, Grid level 3

0.0
0.0

0.1

0.2

0.3

0.4

0.5
y/h

0.6

0.7

0.8

0.9

1.0

Figure 6.18. Axial velocity distribution in section where z/h = 0.2 with the
EARSM on different grid levels.
From figure 6.18 it is concluded that the third grid level was too coarse. The
second grid level was dense enough since the results of the second and
the first grid level are virtually similar.
For demonstrational purposes, also a picture of the cross section of the
quarter of the pipe is presented in figure 6.19. The vorticity in x-direction is
presented as a background colour. Some streamlines are presented in the
foreground with white lines.

62

Test cases

Figure 6.19. Axial vorticity magnitude and some streamlines with the
EARSM.
In the next section the emphasis is diverted to the third order terms.

6.3

Flow inside a rotating pipe

Turbulent flow in pipes has been a popular case for the testing and evaluation of both theories and turbulence models during for many decades, for
example [45], [46] and [47]. According to reference [45], the problem of turbulent flow in an axially rotating pipe has come to be of considerable interest. The laminar counterpart of this problem is not interesting, because the
velocity profile relative to the rotating pipe is identical to its non-rotating
counterpart; i.e. the velocity profile is parabolic. However, for a fully developed turbulent flow there is a profound difference. The flow field is completely three-dimensional yet being a function of only one spatial coordinate
in a cylindrical coordinate system, i.e. radius.

63

Test cases

Relative to an observer, who is rotating with the pipe, the turbulent flow is
no longer unidirectional in the averaged sense. There is a non-zero tangential component of the mean velocity present. Because the continuity
equation forces the radial mean velocity to be zero, there is just one other
component of the mean velocity in addition to the axial component. Furthermore, both these components vary significantly with the rotation rate of
the pipe.

6.3.1 Grid
The grid consisted of a two-degree slice of the pipe because the flow is axisymmetric. The grid was constructed of two blocks. The first block was the
inlet block and consisted of 2 x 160 x 1 cells in the direction of x, y and z,
respectively. The second block, in which the actual solution was calculated,
consisted of 160 x 160 x 1 cells. The grid is presented in figure 6.20.

Figure 6.20. The grid of the rotating pipe. The inlet block is green and the
calculation block red. The figure has been heavily stretched in the ydirection for clarity.
The height of the first block was unity. Its length was 1.6 units. The dimensions of the first cell next to the wall were 0.8 x 0.0005 x 0.035 units at the
outer edge. The height of the second block was unity and its length was
600 units. The second block used the same size of the first cell just at the
beginning of the pipe. The Reynolds number referred to the axial mean velocity and to the pipe diameter was 20 000. The same Reynolds number
was used in the experimental measurements of reference [48], which the
calculations were compared to.

64

Test cases

The non-dimensional y+ value of the lowest cell can be calculated in this


case as

y + = ReD

cf y
2 D

(6.7)

The non-dimensional y-coordinate is therefore y+ = 0.2. The first cells of the


grid inside the boundary layer were thus small enough.

6.3.2 Boundary conditions


The boundary conditions used are presented in table 6.3.
Table 6.3. The boundary conditions for the rotating pipe.
Face
Inlet block
Calculation block

INL

SNG

MIR

CON

MIR

MIR

CON

SNG

PER

EXT

MOV

PER

Where SNG means singular condition meaning that all the fluxes are set
zero. CON means connection between the face numbers four of the inlet
block and one of the calculation block. PER means a periodic boundary
condition between faces three and six of the calculation block. MOV means
a moving wall for which the direction and magnitude are specified in a different file.

6.3.3 Results
The results of the calculations with three different rotation rates are compared to the experimental results of reference [48]. The Reynolds number
based on the axial mean velocity and pipe diameter is 20 000 and the same
Reynolds number was used also in the calculations.

65

Test cases

In figure 6.21 the axial velocity distribution is presented for three different
rotation rates for a fully developed flow.
1.6
1.4
1.2

U(r)/Um

1.0
N = 0, Exp
N = 0, BSL

0.8

N = 0, EARSM
0.6

N = 0.5, Exp
N = 0.5, BSL

0.4

N = 0.5, EARSM
N = 1.0, Exp

0.2

N = 1.0, BSL
N = 1.0, EARSM

0.0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

r/R

Figure 6.21. Axial velocity distributions in a rotating pipe with the EARSM
and the k BSL compared to the experimental results of reference [48].
The tangential velocity distribution is presented in figure 6.22 for two different rotation rates for a fully developed flow.
1.0
N = 0.5, Exp

0.9

N = 0.5, BSL

0.8

N = 0.5, EARSM
0.7
N = 1.0, Exp
V(r)/Vm

0.6

N = 1.0, BSL

0.5

N = 1.0, EARSM

0.4
0.3
0.2
0.1
0.0

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

r/R

Figure 6.22. Tangential velocity distributions in a rotating pipe using the


EARSM and the k BSL compared to the experimental results of reference [48].

66

Test cases

The symbol N in the figures represents the non-dimensional rotation rate,


which is defined as

N=

U (R )
U mean

(6.8)

where U (R ) is the tangential rotation velocity component of the pipe wall


and U mean is the mean axial velocity.
From figure 6.21 it is noted that the k BSL model fails to account the
rotation and gives the same axial velocity profile for each rotation rate
whereas the EARSM is able to follow the trend right when the pipe rotates
faster. However, it underestimates the change in the axial velocity. This is
consistent with the axial velocity profiles and description found from reference [8].
From figure 6.22 it is noted that the k BSL model fails to predict the
right behaviour and gives a solid body rotation for both rotation rates
whereas the EARSM is able to predict the right behaviour. It underestimates the effect of the rotation again and can not make a difference between the different rotation speeds in the tangential velocity profiles,
though. This is also consistent with reference [8]. The prediction of the velocity profiles could be improved by the inclusion of the neglected other cubic term associated with 5 in the tensor polynomial (equation 4.21) and
tuning the coefficients for this special case, as Wallin discusses [8]. It is
worth noticing that the results here are obtained without any tuning whatsoever.
In this case it was also tested whether the solution is grid-independent.
Figure 6.23 represents the results for the axial velocity and figure 6.24 for
the tangential velocity for the case where the non-dimensional rotation rate

N = 0.5.

67

Test cases

1.4

1.2

U(r)/Um

1.0

0.8

0.6

N = 0.5, EARSM, Grid level 1


0.4

N = 0.5, EARSM, Grid level 2


0.2

N = 0.5, EARSM, Grid level 3

0.0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

r/R

Figure 6.23. Axial velocity profiles with different grid levels calculated with
the EARSM.
1.0
0.9

N = 0.5, EARSM, Grid level 1

0.8

N = 0.5, EARSM, Grid level 2

0.7

N = 0.5, EARSM, Grid level 3

V(r)/Vm

0.6
0.5
0.4
0.3
0.2
0.1
0.0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

r/R

Figure 6.24. Tangential velocity profiles with different grid levels calculated
by the EARSM.
From figures 6.23 and 6.24 it is concluded that the third grid level was too
coarse. The second grid level was dense enough since the results do not
change between the second and the first grid level.
The subroutine EARSM1 (see the appendix A) and its correct implementation were also tested using an example found from reference [8]. The mo-

68

Test cases

mentum equation in the tangential direction can be written in the cylindrical


coordinates as

d 2U 1 dU U 2 d (ka r )
ka
+
2 =

+ 2 r
2
r dr
r
dr
r
dr

(6.9)

where U is the local tangential velocity, r the local radial coordinate and

ar the extra anisotropy component. After two integrations, the tangential


velocity can be expressed

U (r ) = U (R )

r r kar

d
R r

(6.10)

where R is the radius of the pipe. According to reference [8], the first term
corresponds to the linear tangential velocity profile and the second term is
the correction that may give a parabolic-like profile if the ar anisotropy is
positive. The results of the test were consistent with this description and the
tangential velocity distribution was exactly as figure 6.22 presents. It is thus
not presented here.
After these different approaches to the test case it is deducted that also the
third order terms of the EARSM are correctly implemented. The EARSM
seems to be working properly with any kind of fully developed flow. The
emphasis is therefore diverted to a developing flow in the last section of this
chapter.

6.4

Flow over a rotating cylinder

The fourth test case was an external axial flow over a cylinder. The first part
of the cylinder does not move and the flow is allowed to develop. At a certain point, where the flow is considered as fully developed, the cylinder
starts to rotate rapidly. This causes a sudden tangential strain into the flow
field and constitutes a relatively simple means of studying threedimensional boundary layer in a developing flow. The results of the calculations are compared to the experimental results made by Bissonette and
Mellor in 1974 [49].

69

Test cases

6.4.1 Grid
The grid consists of a two-degree slice of the cylinder and was constructed
of two blocks. The first block was the inlet block, which consisted of 2 x 192

x 1 cells in the direction of x, y and z, respectively. The second block, in


which the actual solution was calculated, consisted of 224 x 192 x 1 cells.
The grid is presented in figure 6.25.

Figure 6.25. The grid of the rotating cylinder. The inlet block is green and
the calculation block red. The figure has been stretched in the y-direction
for clarity.
The height of the first block was 2.2 units and the length 0.3 units. The grid
was constructed with an inner radius of unity and an outer radius of 3.2
units. The dimensions of the first cell next to the wall were 0.15 x 0.00009 x
0.035 units. The height of the second block was the same as the height of
the first block. The length of the block was 19.4 units. The second block
used the same size of the first cell just at the beginning of the duct. The
Reynolds number referred to the axial mean velocity and to the cylinder radius was 79 500 which is the same as the Reynolds number used in the
experiments of reference [49] that were compared to the calculations. The
non-dimensional y+ value of the lowest cell can be calculated in this case
as

y + = ReR

cf y
2 R

(6.11)

The non-dimensional y-coordinate is therefore y+ = 0.3. The first cells of the


grid inside the boundary layer were thus small enough.
The last part of the cylinder stopped rotating again to simplify the outlet
condition. The parts of the cylinder where the rotation starts and ends can

70

Test cases

be clearly seen in figure 6.25 as the grid in x-direction is more dense


around these areas. The latter part of the cylinder where it suddenly stops
rotating is not studied here. Experimental results for this particular problem
can be found, for example, from reference [50] and CFD results from reference [51].

6.4.2 Boundary conditions


The boundary conditions that were used in the calculations are presented
in table 6.4.
Table 6.4. The boundary conditions for the rotating cylinder.
Face
Inlet block
Calculation block

INL

MIR

MIR

CON

MIR

MIR

CON

MOV

PER

EXT

MIR

PER

Where CON means connection with the face number four of the inlet and
one of the calculation block. PER means periodic boundary condition between faces three and six of the calculation block.

6.4.3 Results
The results are compared to experimental results of reference [49]. They
are drawn at the point were the local Reynolds number based on the mean
flow velocity and x-coordinate from the beginning of the rotating section is
730 000 and x R = 9.1 . The axial and tangential velocity distributions are
presented in figures 6.26 and 6.27.

71

Test cases

1.2

1.0

U/U0

0.8

0.6

0.4
BSL
0.2

EARSM
Experiment

0.0
0.00

0.05

0.10

0.15

0.20

0.25

0.30

0.35

0.40

y/R

Figure 6.26. The axial velocity distribution with the EARSM and the k
BSL compared to the experimental results of reference [49].
1.0
BSL

0.9

EARSM
0.8
Experiment
0.7

W/W0

0.6
0.5
0.4
0.3
0.2
0.1
0.0
0.00

0.05

0.10

0.15

0.20
y/R

0.25

0.30

0.35

0.40

Figure 6.27. The tangential velocity distribution with the EARSM and the
k BSL compared to the experimental results of reference [49].
Surprisingly, the velocity profiles of figures 6.27 and 6.28 are almost identical. Therefore some Reynolds stress distributions are also presented in figures 6.28, 6.29 and 6.30.

72

Test cases

25
BSL
20

EARSM

uu (m/s)2

Experiment
15

10

0
0.00

0.05

0.10

0.15

0.20
y/R

0.25

0.30

0.35

0.40

Figure 6.28. Reynolds stress component u u with the EARSM and the
k BSL compared to the experimental results of reference [49].
8
BSL
7
EARSM
6

Experiment

-uv (m/s)2

5
4
3
2
1
0
0.00

0.05

0.10

0.15

0.20
y/R

0.25

0.30

0.35

0.40

Figure 6.29. Reynolds stress component u v with the EARSM and the
k BSL compared to the experimental results of reference [49].

73

Test cases

6
BSL
5

EARSM
Experiment

vw (m/s)2

0
0.00

0.05

0.10

0.15

0.20

0.25

0.30

0.35

0.40

y/R

Figure 6.30. Reynolds stress component vw with the EARSM and the
k BSL compared to the experimental results of reference [49].
For demonstrational purposes, also a picture of the same Reynolds stress
component vw as in figure 6.30 is shown in figure 6.31 as this is the Reynolds stress component that corresponds with the tangential shear stress
component. The figure is calculated with the EARSM and shows an area of
the calculation space close to the cylinder surface. The points where the
rotation starts and stops are clearly seen in the figure.

Figure 6.31. The Reynolds stress component vw with the EARSM close
to the cylinder surface.

74

Test cases

The local skin friction coefficients are presented in figure 6.32 for the axial
friction and in figure 6.33 for the tangential friction. The local Reynolds
numbers are referred to the axial mean velocity and the distance from the
beginning of the rotating section.
0.008

0.007

0.006

Cf

0.005

0.004

0.003

Axial, BSL

0.002

Axial, EARSM
0.001

0.000
-2.0E+05

Axial, Exp

-1.0E+05

0.0E+00

1.0E+05

2.0E+05

3.0E+05

4.0E+05

5.0E+05

6.0E+05

Local Reynolds number

Figure 6.32. The axial friction coefficients with the EARSM and the k
BSL compared to the experimental results of reference [49].
0.008

0.007

0.006

Cf

0.005

0.004

0.003

Tangential, BSL
0.002

Tangential, EARSM
0.001

0.000
-2.0E+05

Tangential, Exp

-1.0E+05

0.0E+00

1.0E+05

2.0E+05

3.0E+05

4.0E+05

5.0E+05

6.0E+05

Local Reynolds number

Figure 6.33. The tangential friction coefficients with the EARSM and the
k BSL compared to the experimental results of reference [49].

75

Test cases

As the last calculation, the grid-independence of the solution was tested.


Figure 6.34 represents the results for the axial velocity using different grid
levels. The results are from the point where the local Reynolds number
based on the mean flow velocity and distance from the beginning of the
rotation is 730 000.
1.2

1.0

U/U0

0.8

0.6

EARSM, Grid level 1

0.4

EARSM, Grid level 2


0.2
EARSM, Grid level 3
0.0
0.00

0.05

0.10

0.15

0.20

0.25

0.30

0.35

0.40

y/R

Figure 6.34. Axial velocity profiles with different grid levels calculated with
the EARSM.
From figure 6.34 it is concluded that the third grid level was already dense
enough since the results do not change between the third and the second
grid level. The calculation was performed on the second grid level to make
sure that also the skin friction coefficients were also correct. The calculation
on the first grid level was useless in the terms of the computer resources.
Based on the results of this last test case it is concluded that the EARSM is
working properly also in the case of a developing three dimensional flow.
However, the results of the EARSM and k BSL are very similar and no
large differences are noted. This can not be seen as a shortcoming of the
EARSM mainly because the results of the k BSL model are already
very good. Still, it was expected that the results of the EARSM would differ
more from the corresponding results of the k BSL. The following discussion tries to explain why the author assumed such an expectation.

76

Test cases

If a new tangential strain is introduced to the EARSM, the absolute values


of both the 1 - and 6 -coefficients reduce. Both of the coefficients being
always negative the absolute value of the 1 + II 6 -coefficient reduces as
well. Therefore it was expected that the absent 5 -coefficient, or the corresponding third order term, would account for the new tangential strain,
which should increase the axial shear stress and the skin friction coefficient
as well. The invariants can not be responsible for such an increase because, by definition, they are directionally invariant.
From the results of the fourth test case it is concluded that the EARSM
does not seem to take the sudden tangential strain directly into account.
However, an indirect effect is perceived. The tangential strain causes the
production of turbulence to increase, which in turn increases the turbulent
kinetic energy. All the effects of the turbulence increase, such as the shear
stress and friction coefficients. However, this process is somewhat slow
and can not precisely model the sudden additional strain at the beginning of
the rotation as the results reveal.
It can also be noted that the turbulent kinetic energy does not grow fast
enough in the overall level because the EARSM can not model accurately
the streamline curvature. This is caused by the simplification made during
the formulation of the ARSM, particularly the neglecting of the convection
term due to the equilibrium approximation. The approximation is made in
the Cartesian coordinate system and cannot reproduce strong streamline
curvature exactly. The equilibrium approximation should be made in a
streamline coordinate system instead, but it leads to an extremely complicated model. The convection term of the Reynolds stress anisotropy tensor
should perhaps be approximated somehow. At the moment of writing of this
text, better formulations have not yet been published. However, it must
mentioned that there have been a few attempts, such as references [52]
and [53], but they are not very practical, yet.

77

Chapter 7
Summary and conclusions

The goal of the project was to implement an explicit algebraic Reynolds


stress model of reference [8] into the Navier-Stokes flow solver FINFLO.
First the Navier-Stokes equations, a brief description of turbulence and the
eddy-viscosity turbulence models were presented. Then the differential
Reynolds stress model was described briefly and the algebraic Reynolds
stress model as well as its explicit form were formulated. Then some programming aspects of the EARSM were discussed. At the end four test
cases were selected, calculated and analysed to verify the correct implementation of the model.
As the first test case the two-dimensional boundary layer over a flat plate
was calculated. This verified that the eddy-viscosity coefficient of the
EARSM was correctly implemented. As the second test case a fully developed flow inside a rectangular duct was calculated to verify the correct implementation of the second order terms. They were also concluded as implemented correctly, as the secondary flow patterns appeared. As the third
test case a fully developed flow inside a rotating pipe was calculated. The
velocity profiles indicated that also the third order terms of the model were
correctly implemented. As the last test case an axial flow over a cylinder,
part of which rotates, was calculated to further study the behaviour of the

79

Summary and conclusions

EARSM under the application of a sudden strain component. This test case
also proved to be successful and increased the understanding in the physical and numerical behaviour of the EARSM.
It was also concluded that the second and third order terms of the EARSM
are usually of only minor importance in describing the flow field. Their effect
remains fairly small in most cases of engineering interest. However, there is
one major advantage in the formulation of the EARSM over the eddyviscosity models. The effective eddy-viscosity coefficient of the EARSM
adapts to the local strain and vorticity field by their invariants. In the linear
eddy-viscosity models there is no such eddy-viscosity coefficient, i.e. the
coefficient can be thought of as being constant. Actually, when the eddyviscosity coefficient of the EARSM is assigned a value of unity, the model
behaves similarly to the k BSL.
To sum the thesis up, it can be said that EARSM has proved to be an improvement of the flow solver FINFLO. The author hopes it will be used in
many practical applications in future.

80

Bibliography

[1]

D. A. Anderson, J. C. Tannehill and Richard H. Pletcher, Computa-

tional Fluid Mechanics and Heat Transfer, Hemisphere Publishing


Corporation, U.S.A., 1984, ISBN 0-89116-471-5.
[2]

B. S. Baldwin and H. Lomax, Thin Layer Approximation and Alge-

braic Model for Separated Turbulent Flows, AIAA Paper, 1978, pp.
78-257.
[3]

H. W. Stock and W. Haase, Determination of Length Scales in Al-

gebraic Turbulence Models for Navier-Stokes Methods , AIAA Journal, Vol. 27, No. 1, 1989, pp. 5-14.
[4]

K.Y. Chien, Predictions of Channel and Boundary-Layer Flows

with a Low-Reynolds-Number Turbulence Model , AIAA Journal, Vol.


20, No. 1, 1982, pp. 33-38.
[5]

F. R. Menter, Two-Equation Eddy-Viscosity Turbulence Model for

Engineering Applications, AIAA Journal, Vol. 32, 1994, pp. 15981605.


[6]

F. R. Menter, Zonal Two Equation k Turbulence Models for

Aerodynamic Flows, In 24th AIAA Fluid Dynamics Conference, Orlando, Florida, U.S.A., 1993, AIAA Paper 93-2906.

81

Bibliography

[7]

A. Hellsten, Some Improvements in Menters k SST Turbulence

Model, In 29th AIAA Fluid Dynamics Conference, Albuquerque, New


Mexico, U.S.A., 1998, AIAA paper 98-2554.
[8]

S. Wallin, Engineering turbulence modeling for CFD with a focus on

explicit algebraic Reynolds stress models , Doctoral Thesis,


Norsteds Tryckeri AB, Stockholm, Sweden, 2000.
[9]

P. Kaurinkoski and A. Hellsten, FINFLO: The Parallel Multi-Block

Flow Solver, Report A-17, Laboratory of Aerodynamics, Helsinki


University of Technology, Espoo, Finland, 1998.
[10]

P. Kaurinkoski, Development of an Equation of State for an Arbitrary

Mixture of Thermally Perfect Gases to the FINFLO flow solver , Report No B-48, Series B, Laboratory of Aerodynamics, Helsinki University of Technology, Finland, 1995.
[11]

P. Kaurinkoski and A. Hellsten, FINFLO Users Guide, Version 5.1,


Laboratory of Aerodynamics, Helsinki University of Technology,
Finland, 1998, (unpublished).

[12]

T. Brandt, Inclusion of a Two-Equation Turbulence Model in the

Navier-Stokes Solver FINFLO-SHIP, Masters Thesis, Helsinki University of Technology, Finland, 2000.
[13]

J. Hoffren and T. Siikonen, FINF2A: a Multi-Block Navier-Stokes

Solver for Steady Two-Symmetric and Axisymmetric flows, Report


No B-38, Series B, Laboratory of Aerodynamics, Helsinki University
of Technology, Finland, 1992.
[14]

D. C. Wilcox, Turbulence Modeling for CFD, second edition, DCW


Industries, Inc., La Canada, U.S.A., 1998, ISBN 0-9636051-5-1.

[15]

B. A. Younis, Applied Turbulence modelling, Lecture Notes, Helsinki


University of Technology, Finland, 1998, (unpublished).

[16]

P. R. Spalart and S. R. Allmaras, A One-Equation Turbulence

Model for Aerodynamic Flows, AIAA Fluid Dynamics Conference,


Reno, Nevada, U.S.A, 1992, AIAA paper 92-439.

82

Bibliography

[17]

B. S. Baldwin and T. J. Barth, A One-Equation Turbulence Trans-

port Model for High Reynolds Number Wall-Bounded Flows, NASA


TM-102847, 1990.
[18]

A. Hellsten, Implementation of a One-Equation Turbulence Model

into the FINFLO Flow Solver, Report B-49, Series B, Laboratory of


Aerodynamics, Helsinki University of Technology, Finland, 1996.
[19]

W. P. Jones and B. E. Launder, The Prediction of Laminarization

with a Two-Equation Model of Turbulence, Int. J. Heat and Mass


Transfer, Vol. 15, 1972, pp. 301-314.
[20]

M. A. Leschziner, J. J. McGuirk, et.al., Turbulence modeling for

CFD applications course lecture material , Department of Mechanical engineering, UMIST, UK, 1996, (unpublished).
[21]

B. J. Daly and E. H. Harlow, Transport Equations in Turbulence ,


Phys. Fluids, Vol. 13, 1970, pp. 2634-2649.

[22]

K. Hanjalic and B. E. Launder, A Reynolds Stress Model of Turbu-

lence and Its Application to Thin Shear Flows, J. Fluid Mech., Vol.
52, 1972, pp. 609-.
[23]

J. L. Lumley and B. Khajeh-Nouri, Computational Modelling of Tur-

bulent Transport, Adv. Geophys., Vol. 8A, 1974, pp. 169-.


[24]

B, E. Launder, G. J. Reece and W. Rodi, Progress in the Develop-

ment of a Reynolds Stress Turbulence Closure, J. Fluid Mech., Vol.


41, 1975, pp. 537-566.
[25]

C. G. Speziale, S. Sarkar and T. B. Gatski, Modelling the Pressure-

Strain Correlation of Turbulence: an Invariant Dynamical Systems


Approach, J. Fluid Mech., Vol. 227, 1991, pp. 245-272.
[26]

S. S. Girimaji, Fully Explicit and Self-Consistent Algebraic Reynolds

Stress Model, Theoretical and Computational Fluid Dynamics, Vol.


8, 1996, pp. 387-402.

83

Bibliography

[27]

W. Rodi, The prediction of free turbulent boundary layers by use of

a two equation model of turbulence, Ph.D. Thesis, University of


London, 1972.
[28]

T. Siikonen and H. Pan, Application of Roes Method for the Simu-

lation of Viscous Flow in Turbomachinery, Proceedings of the first


European Computational Fluid Dynamics Conference, Brussels,
Belgium, 1992, Elsevier Science Publishers B.V., pp. 635-641.
[29]

J. C. Rotta, Statistische Theorie Nichthomogener Turbulenz, Z.


Phys., Vol. 129, 1951, pp. 547-572.

[30]

S. B. Pope, A More General Effective-Viscosity Hypothesis, J. of


Fluid Mech., Vol. 72, 1975, pp. 331-340.

[31]

D. B. Taulbee, An Improved Algebraic Reynolds Stress Model and

Corresponding Nonlinear Stress Model, Physics of Fluids A, Vol. 4,


1992, pp. 2555-2560.
[32]

D. B. Taulbee, J. R. Sonnenmeier and K. M. Wall, Stress Relation

for Three-Dimensional Turbulent Flows, Physics of Fluids, Vol. 6,


1994, pp. 1399-1401.
[33]

T. B. Gatski and C. G. Speziale, On Explicit Algebraic Stress Mod-

els for Complex Turbulent Flows, J. of Fluid Mech., Vol. 254, 1993,
pp. 59-78.
[34]

J. L. Lumley, Computational Modeling of Turbulent Flows, Adv.


Appl. Mech., Vol. 18, 1978, pp. 123-176.

[35]

A. Shabbir and T. H. Shih, Critical Assessment of Reynolds Stress

Turbulence Models Using Homogeneous Flows, NASA TM 105954,


ICOMP-92-24, CMOTT-92-12, 1992.
[36]

A. Hellsten and S. Laine, Explicit Algebraic Reynolds-Stress Model-

ling in Decelerating and Separating Flows, In AIAA Fluid Dynamics


Conference, Denver, Colorado, U.S.A., 2000, AIAA paper 20002313.

84

Bibliography

[37]

A. Hellsten and P. Rautaheimo (editors), 8th ERCOFTAC/IAHR/

COST Workshop on Refined Turbulence Modelling, Report 127,


Laboratory of Applied Thermodynamics, Helsinki University of
Technology, 1999.
[38]

P. S. Klebanoff, Characteristics of Turbulence in a Boundary Layer

with Zero Pressure Gradient , Report 1247, National Advisory Committee for Aeronautics (NACA), U.S.A., 1955.
[39]

P. J. Roache, Fundamentals of Computational Fluid Dynamics,


Hermosa Publishers, U.S.A., 1998, ISBN 0-913478-09-1.

[40]

K. G. Winter and L. Gaudet, Turbulent Boundary Layer Studies at

High Reynolds Numbers at Mach Numbers between 0.2 and 2.8,


Aeronautical Research Council, Reports and Memoranda, R. & M.
No. 3712, London, UK, 1973.
[41]

F. M. White, Viscous Fluid Flow, second edition, McGraw-Hill Book


Co., Singapore, 1991, ISBN 0-07-100995-7.

[42]

B. A. Petterson Reif and H. I. Andersson, Second-moment closure

predictions of turbulence-induced secondary flow in a straight


square duct, Engineering Turbulence Modelling and Experiments 4,
Elsevier Science Ltd., The Netherlands, 1999, ISBN 0-08-043328-6,
pp. 349-358.
[43]

T. Rung, H. Lbcke and F. Thiele, Turbulence Closure Model Con-

straint Derived from Stress-Induced Secondary Flow, AIAA Journal,


Vol. 38, No. 9, 2000, pp. 1756-1758.
[44]

H. Yokosawa, H. Fujita, M. Hirota and S. Iwata, Measurement of

turbulent flow in a square duct with roughened walls on two opposite sides, Int. J. Heat and Fluid Flow, Vol. 10, No. 2, 1989, pp. 125130.
[45]

C. G. Speziale, B. A. Younis and S. A. Berger, Analysis and model-

ling of turbulent flow in an axially rotating pipe, J. Fluid Mech., Vol.


407, 2000, pp. 1-26.

85

Bibliography

[46]

Z. Yang, Large eddy simulation of fully developed turbulent flow in a

rotating pipe, Int. J. Numer. Meth. Fluids, Vol. 33, 2000, pp. 681694.
[47]

B. A. Pettersson, H. I. Andersson and A. S. Brunvoll, Modelling

Near-Wall Effects in Axially Rotating Pipe Flow by Elliptic Relaxation, AIAA Journal, Vol. 36, No. 7, 1998, pp. 1164-1170.
[48]

S. Imao, M. Itoh and T. Harada, Turbulent characteristics of the flow

in an axially rotating pipe, Int. J. Heat and Fluid Flow, Vol. 17, No. 5,
1996, pp. 444-451.
[49]

L. R. Bissonette and G. L. Mellor, Experiment of the behaviour of an

axisymmetric turbulent boundary layer with a sudden circumferential


strain, J. Fluid Mech., Vol. 63, Part 2, 1974, pp. 369-413.
[50]

D. M. Driver and S. K. Hebbar, Experimental Study of a Three-

Dimensional, Shear-Driven, Turbulent Boundary Layer, AIAA Journal, Vol. 25, No. 1, 1987, pp. 35-42.
[51]

I. Bassina, M. Strelets and P. R. Spalart, Response of Simple Tur-

bulence Models to Step Changes of Slip Velocity, AIAA Journal,


Vol. 39, No. 2, 2001, pp. 201-210.
[52]

S. S. Girimaji, A Galilean invariant explicit algebraic Reynolds stress

model for turbulent curved flows , Phys. Fluids, Vol. 9, 1997, pp.
1067-1077.
[53]

C. L. Rumsey, T. L. Gatski and J. H. Morrison, Turbulence model

predictions of extra-strain rate effects in strongly-curved flows , AIAA


paper 99-157, 1999.

86

Appendix A, Subroutine EARSM1


C

SUBROUTINE EARSM1(VIS,EPS2,EPS4,RK,DUIDXJ,BIJ,IMAX,JMAX,KMAX,TTS)
*****************

C
C
C

A subroutine to calculate the extra anisotropy components and


the effective eddy-viscosity coefficient used with the EARSM
23.11.2001 Ville H. / Laboratory of Aerodynamics
USE MAIN_ARRAYS , ONLY : MAXEB
USE CONSTANTS , ONLY : IN,JN,KN
USE CASE_DEFS , ONLY : PR,PRT,TURLIM,FRSMUT
IMPLICIT NONE
REAL , INTENT(IN) , DIMENSION(*) :: VIS,RK,TTS
REAL , DIMENSION(*) :: EPS2,EPS4
REAL , INTENT(IN) , DIMENSION(MAXEB,9) :: DUIDXJ
REAL , DIMENSION(MAXEB,5) :: BIJ
INTEGER , INTENT(IN) :: IMAX,JMAX,KMAX
REAL ::
INTEGER
REAL ::
REAL ::
REAL ::
REAL ::
REAL ::
REAL ::
REAL ::
REAL ::
REAL ::

P3,P6,TT,C1,C1P,C1P3,C1PSQ,PRS,EPS2UP,EPS4UP
:: ISTRID,JSTRID,IL,KA,K,KK,J,JJ,I,L
S11,S22,S33,S12,S13,S23,W12,W13,W23
S11SQ,S22SQ,S33SQ,S12SQ,S13SQ,S23SQ,W12SQ,W13SQ,W23SQ
HTTS,SII,WII,WIIP3,SWWIV,SWWIVTT,SSWWV
TERM3C11,TERM3C12,TERM3C13,TERM3C22,TERM3C23
TERM4C11,TERM4C12,TERM4C13,TERM4C22,TERM4C23
TERM6C11,TERM6C12,TERM6C13,TERM6C22,TERM6C23
TERM9C11,TERM9C12,TERM9C13,TERM9C22,TERM9C23
P1,P2,SQP2,PM,PMP,SQPM,FACOS,FNC,D,FII1,FII2,FN,FNSQ,Q,PQ
BETA1,BETA3,BETA4,BETA6,BETA9,CMUEFF,PVIS,EPS2LO,EPS4LO

Often used numbers


P3
= 1.0/3.0
P6
= 0.5*P3
TT
= 2.0*P3

! One third
! One sixth
! Two thirds

Model coefficient and its variants


C1
= 1.8
C1P
= 9.0*(C1 - 1.0)/4.0 ! C1_prime
C1P3
= C1P*P3
C1PSQ = C1P**2

Upper limits of EPS2 and EPS4


PRS
= PR/PRT
EPS2UP = 1.0 + TURLIM
EPS4UP = 1.0 + TURLIM*PRS

Constants for the loops over the grid


ISTRID = IMAX + 2*IN
! Total number of cells in i-direction
JSTRID = JMAX + 2*JN
! Total number of cells in j-direction
IL
= ISTRID*JSTRID
! Total number of cells in ij-plane
KA
= (KN-1)*IL + (JN-1)*ISTRID + IN ! The start address

Actual loops
DO K = 1, KMAX
KK = K*IL + KA
DO J = 1, JMAX
JJ = J*ISTRID + KK
DO I = 1, IMAX
L = JJ + I

Half of the time scale


HTTS
= 0.5*TTS(L)

Strain-rate and vorticity tensor components


S11
= TTS(L)*DUIDXJ(L,1)

87

Appendix A, Subroutine EARSM1

S22
S33

= TTS(L)*DUIDXJ(L,5)
= TTS(L)*DUIDXJ(L,9)

S12
S13
S23

= HTTS*(DUIDXJ(L,2) + DUIDXJ(L,4))
= HTTS*(DUIDXJ(L,3) + DUIDXJ(L,7))
= HTTS*(DUIDXJ(L,6) + DUIDXJ(L,8))

W12
W13
W23

= HTTS*(DUIDXJ(L,2) - DUIDXJ(L,4))
= HTTS*(DUIDXJ(L,3) - DUIDXJ(L,7))
= HTTS*(DUIDXJ(L,6) - DUIDXJ(L,8))

Squares of strain-rate and vorticity tensor components


S11SQ = S11*S11
S22SQ = S22*S22
S33SQ = S33*S33
S12SQ
S13SQ
S23SQ

= S12*S12
= S13*S13
= S23*S23

W12SQ
W13SQ
W23SQ

= W12*W12
= W13*W13
= W23*W23

Second invariants of the strain rate and vorticity tensors


SII
= S11SQ + S22SQ + S33SQ + 2.0*(S12SQ+S13SQ+S23SQ)
WII
=-2.0*(W12SQ + W13SQ + W23SQ)
WIIP3 = P3*WII ! One third of the invariant

Third invanriant of the strain rate and vorticity tensors:


SWWIV =-S11*(W12SQ + W13SQ) - S22*(W12SQ + W23SQ)
&
- S33*(W13SQ + W23SQ)
&
+ 2.0*(-S12*W13*W23 + S13*W12*W23 - S23*W12*W13)
SWWIVTT= TT*SWWIV ! Two thirds of the invariant

Fourth invariant of the strain rate and vorticity tensors


SSWWV = 2.0*(-(S12*S13 + S22*S23 + S23*S33)*W12*W13
&
+(S11*S13 + S12*S23 + S13*S33)*W12*W23
&
-(S11*S12 + S12*S22 + S13*S23)*W13*W23)
&
- (S11SQ + S12SQ + S13SQ)*(W12SQ + W13SQ)
&
- (S12SQ + S22SQ + S23SQ)*(W12SQ + W23SQ)
&
- (S13SQ + S23SQ + S33SQ)*(W13SQ + W23SQ)

Tensor component terms for beta 3


TERM3C11 = - W12SQ - W13SQ - WIIP3
TERM3C22 = - W12SQ - W23SQ - WIIP3
TERM3C12 = - W13*W23
TERM3C13 =
W12*W23
TERM3C23 = - W12*W13

Tensor component terms for beta 4


TERM4C11 =-2.0*(S12*W12 + S13*W13)
TERM4C22 = 2.0*(S12*W12 - S23*W23)
TERM4C12 = (S11-S22)*W12
- S23*W13
- S13*W23
TERM4C13 =
- S23*W12 + (S11-S33)*W13
+ S12*W23
TERM4C23 =
S13*W12
+ S12*W13 + (S22-S33)*W23

tensor component terms for beta 6


TERM6C11 = -2.0*((S12*W13 - S13*W12)*W23
&
+ S11*(W12SQ + W13SQ)) - SWWIVTT
TERM6C22 = -2.0*((S23*W12 + S12*W23)*W13
&
+ S22*(W12SQ + W23SQ)) - SWWIVTT
TERM6C12 = -S12*(2.0*W12SQ + W13SQ + W23SQ)
&
- (S13*W13-S23*W23)*W12 - (S11+S22)*W13*W23
TERM6C13 = -S13*(W12SQ + 2.0*W13SQ + W23SQ)
&
- (S12*W12+S23*W23)*W13 + (S11+S33)*W12*W23
TERM6C23 = -S23*(W12SQ + W13SQ + 2.0*W23SQ)
&
+ (S12*W12-S13*W13)*W23 - (S22+S33)*W12*W13

88

- WII*S11
- WII*S22
- WII*S12
- WII*S13
- WII*S23

Appendix A, Subroutine EARSM1

Tensor component terms for beta 9


TERM9C11 =-2.0*(( S12*W12 + S13*W13 - S23*W23)*W12SQ
&
+( S12*W12 + S13*W13 + S23*W23)*W13SQ
&
+( S22-S33)*W12*W13*W23)
TERM9C22 =-2.0*((-S12*W12 - S13*W13 + S23*W23)*W12SQ
&
+(-S12*W12 + S13*W13 + S23*W23)*W23SQ
&
+(-S11+S33)*W12*W13*W23)
TERM9C12 = ((S11-S22)*W12 - 2.0*(S13*W23+S23*W13))*W12SQ
&
+ ((S11-S33)*W12 - 2.0* S13*W23)
*W13SQ
&
+ ((S33-S22)*W12 - 2.0* S23*W13)
*W23SQ
TERM9C13 = ((S11-S22)*W13 + 2.0* S12*W23)
*W12SQ
&
+ ((S11-S33)*W13 + 2.0*(S12*W23-S23*W12))*W13SQ
&
+ ((S22-S33)*W13 - 2.0* S23*W12)
*W23SQ
TERM9C23 = ((S22-S11)*W23 + 2.0* S12*W13)
*W12SQ
&
+ ((S11-S33)*W23 + 2.0* S13*W12)
*W13SQ
&
+ ((S22-S33)*W23 + 2.0*(S12*W13+S13*W12))*W23SQ

Solution of the third degree equation for N_c


P1
= (C1PSQ/27.0 + 0.45*SII - TT*WII)*C1P
P2
= P1**2 - (C1PSQ*P3**2 + 0.9*SII + TT*WII)**3
IF (P2 .GE. 0.0) THEN
SQP2 = SQRT(P2)
PM
= P1 - SQP2
PMP
= ABS(PM)**P3
FNC
= C1P3 + (P1+SQP2)**P3 + SIGN(PMP,PM)
ELSE
PM
= P1**2 - P2
SQPM = SQRT(PM)
FACOS = P3*ACOS(P1/SQPM)
FNC
= C1P3 + 2.0*(PM**P6)*COS(FACOS)
END IF

Improvement of the approximation of the N


D
= 20.0*(FNC**4)*(FNC - 0.5*C1P)
&
- WII*(10.0*FNC + 15.0*C1P)*FNC**2
&
+ 10.0*C1P*WII**2
FII1
= SWWIV**2
FII2
= SSWWV - 0.5*SII*WII
FN
= FNC + 162.0*(FII1 + FII2*FNC**2)/D

The denominator of the betas


FNSQ
= FN**2
Q
= 5.0*P6*(FNSQ - 2.0*WII)*(2.0*FNSQ - WII)
PQ
= 1.0/Q

Coefficients (betas)
BETA1
= BETA3
= BETA4
= BETA6
= BETA9
=

C
C

PQ*FN*(2.0*FNSQ - 7.0*WII)
PQ*12.0*SWWIV/FN
PQ*2.0*(FNSQ - 2.0*WII)
PQ*6.0*FN
PQ*6.0

Extra anisotropy components. Note that we use b_ij and Wallin


uses a_ij, which is a_ij = 2*b_ij.
BIJ(L,1) = 0.5*(BETA3*TERM3C11 + BETA4*TERM4C11
&
+ BETA6*TERM6C11 + BETA9*TERM9C11)
BIJ(L,2) = 0.5*(BETA3*TERM3C12 + BETA4*TERM4C12
&
+ BETA6*TERM6C12 + BETA9*TERM9C12)
BIJ(L,3) = 0.5*(BETA3*TERM3C13 + BETA4*TERM4C13
&
+ BETA6*TERM6C13 + BETA9*TERM9C13)
BIJ(L,4) = 0.5*(BETA3*TERM3C22 + BETA4*TERM4C22
&
+ BETA6*TERM6C22 + BETA9*TERM9C22)
BIJ(L,5) = 0.5*(BETA3*TERM3C23 + BETA4*TERM4C23
&
+ BETA6*TERM6C23 + BETA9*TERM9C23)
Eps2 and Eps4
PVIS
CMUEFF
EPS2(L)
EPS4(L)

=
=
=
=

1.0/VIS(L)
-0.5*(BETA1 + WII*BETA6)
1.0 + CMUEFF*RK(L)*TTS(L)*PVIS
1.0 + CMUEFF*RK(L)*TTS(L)*PVIS*PRS

89

Appendix A, Subroutine EARSM1

Lo-limiter (minimum)
EPS2LO = 1.0 + FRSMUT*PVIS
EPS4LO = 1.0 + FRSMUT*PVIS*PRS
EPS2(L) = MAX(EPS2LO, EPS2(L))
EPS4(L) = MAX(EPS4LO, EPS4(L))

Hi-limiter (maximum)
EPS2(L) = MIN(EPS2UP, EPS2(L))
EPS4(L) = MIN(EPS4UP, EPS4(L))
END DO
END DO
END DO

! End of I-loop
! End of J-loop
! End of K-loop

RETURN
END

90

Appendix B, Subroutine TTIMES


C

SUBROUTINE TTIMES(RO,VIS,RK,REPS,ITURB,IMAX,JMAX,KMAX,TTS)
*****************

C
C
C

A subroutine to calculate the turbulent time scale for


the EARSM
2.11.2000 Ville H. / Laboratory of Aerodynamics
USE CONSTANTS , ONLY : IN,JN,KN
IMPLICIT NONE
INTEGER , INTENT(IN) :: ITURB,IMAX,JMAX,KMAX
INTEGER :: ISTRID,JSTRID,IL,KA,K,KK,J,JJ,K,L,I
REAL , INTENT(IN) , DIMENSION(*) :: VIS,RO,RK,REPS
REAL , DIMENSION(*) :: TTS
REAL :: CT,BSTAR,PREPS,TSTUR,TSVIS,BNUM

Model coefficients
CT
= 6.0
BSTAR = 0.09

Constants for the loops over the grid


ISTRID = IMAX + 2*IN
! Total number of cells in i-direction
JSTRID = JMAX + 2*JN
! Total number of cells in j-direction
IL
= ISTRID*JSTRID
! Total number of cells in ij-plane
KA
= (KN-1)*IL + (JN-1)*ISTRID + IN ! The start address
IF (ITURB.EQ.3) THEN
! k-epsilon models
DO K = 1, KMAX
KK = K*IL + KA
DO J = 1, JMAX
JJ = J*ISTRID + KK
DO I = 1, IMAX
L = JJ + I
PREPS = 1.0/REPS(L)
TSTUR = RK(L)*PREPS
TSVIS = CT*SQRT(VIS(L)*PREPS)
TTS(L) = MAX(TSTUR,TSVIS)
END DO
END DO
END DO
ELSE IF (ITURB.EQ.6) THEN ! k-omega models
DO K = 1, KMAX
KK = K*IL + KA
DO J = 1, JMAX
JJ = J*ISTRID + KK
DO I = 1, IMAX
L = JJ + I
BNUM
= 1.0/(BSTAR*REPS(L))
TSTUR = RO(L)*BNUM
TSVIS = CT*SQRT(RO(L)*VIS(L)*BNUM/RK(L))
TTS(L) = MAX(TSTUR,TSVIS)
END DO
END DO
END DO

C
C

ELSE IF (ITURB.EQ.X) THEN ! k-tau models


k-tau models are not in use, yet
END IF
RETURN
END

91

Appendix C, Subroutine ANISFLUX

SUBROUTINE ANISFLUX(IMAX,JMAX,KMAX,JADD,KADD,IL,U,V,W,RK,
&
BIJ,A,AX,AY,AZ,FRM,FRN,FRW,FE)
*******************

C
C
C

A subroutine to add the anisotropic parts of the Reynolds


stresses to the momentum and energy fluxes
13.11.2000 Ville H. / Laboratory of Aerodynamics
USE MAIN_ARRAYS , ONLY : MAXEB
USE CONSTANTS , ONLY : IN,JN,KN
IMPLICIT NONE
INTEGER , INTENT(IN) :: IMAX,JMAX,KMAX,JADD,KADD,IL
REAL , INTENT(IN) , DIMENSION(*) :: U,V,W,RK,A,AX,AY,AZ
REAL , INTENT(IN) , DIMENSION(MAXEB,5) :: BIJ
REAL , DIMENSION(*) :: FRM,FRN,FRW,FE
INTEGER :: ISTRID,JSTRID,LSTRID,IJSLB,KG,IA,IG,I,IMIL
REAL :: UCM,VCM,WCM,B11CM,B12CM,B13CM,B22CM,B23CM,B33CM
REAL :: RKA,F2EX,F3EX,F4EX,F5EX

Constants for the loops over the grid


ISTRID = IMAX + 2*IN
! Total number of cells in i-direction
JSTRID = JMAX + 2*JN
! Total number of cells in j-direction
LSTRID = ISTRID*JSTRID
! Total number of cells in ij-plane
IJSLB = (JMAX+JADD)*ISTRID ! Number of cells in ij-slab
DO KG = 1,KMAX+KADD
IA = (KG+KN-1)*LSTRID + JN*ISTRID
DO IG = 1,IJSLB
I
= IA + IG
IMIL
= I - IL
! IL = ISTR, JSTR or KSTR

Mean values of velocity components


UCM
= 0.5*(U(I) + U(IMIL))
VCM
= 0.5*(V(I) + V(IMIL))
WCM
= 0.5*(W(I) + W(IMIL))

C
C

Twice the mean values of extra anisotrophy components b_ij


Remember that a_ij = 2.0*b_ij
B11CM = BIJ(I,1) + BIJ(IMIL,1)
B12CM = BIJ(I,2) + BIJ(IMIL,2)
B13CM = BIJ(I,3) + BIJ(IMIL,3)
B22CM = BIJ(I,4) + BIJ(IMIL,4)
B23CM = BIJ(I,5) + BIJ(IMIL,5)
B33CM =-BIJ(I,1) - BIJ(I,4) - BIJ(IMIL,1) - BIJ(IMIL,4)

The change of the momentum and energy fluxes


RKA
= 0.5*(RK(I) + RK(IMIL))*A(I)
F2EX
= RKA*(AX(I)*B11CM + AY(I)*B12CM + AZ(I)*B13CM)
F3EX
= RKA*(AX(I)*B12CM + AY(I)*B22CM + AZ(I)*B23CM)
F4EX
= RKA*(AX(I)*B13CM + AY(I)*B23CM + AZ(I)*B33CM)
F5EX
= UCM*F2EX + VCM*F3EX + WCM*F4EX

Updated momentum and energy fluxes


FRM(I) = FRM(I) + F2EX
FRN(I) = FRN(I) + F3EX
FRW(I) = FRW(I) + F4EX
FE(I) = FE(I) + F5EX
END DO
END DO
RETURN
END

93

Appendix D, Subroutine FLUXP

SUBROUTINE FLUXP(FRM,FRN,FRW,RO,RM,RN,RW,P,RK,BIJ,UROT,A,AX,AY,AZ,
&
IMAX,JMAX,KMAX,IL,JADD,KADD)
****************

C
C
C

A subroutine to calculate the momentum fluxes for an axisymmetric


flow-case. Increment in DIM=1 direction is IL.
Modified 26.2.2001 Ville H. / Laboratory of Aerodynamics
USE MAIN_ARRAYS , ONLY : MAXEB
USE CONSTANTS , ONLY : IN,JN,KN
USE CASE_DEFS
IMPLICIT NONE
REAL , DIMENSION(*) :: FRM,FRN,FRW
REAL , INTENT(IN) , DIMENSION(*) :: RO,RM,RN,RW,P,RK
REAL , INTENT(IN) , DIMENSION(*) :: A,AX,AY,AZ,UROT
INTEGER , INTENT(IN) :: IMAX,JMAX,KMAX,IL,JADD,KADD
REAL , PARAMETER :: P23 = 2.0/3.0
REAL :: PAM,UM,VM,WM,VFLOM,PERRO,RMCM
REAL :: B11CM,B12CM,B13CM,B22CM,B23CM,B33CM,RKA,F2EX,F3EX,F4EX
REAL , INTENT(IN) , DIMENSION(MAXEB,5) :: BIJ
INTEGER :: II,IM,ISTRID,JSTRID,IJSLB,KG,IG,IA
JSTRID = JMAX + 2*JN
ISTRID = IMAX + 2*IN
IJSLB

= ISTRID*(JMAX+JADD) - IN + 1

IF (ITURB.GE.3) THEN
! 2-eq models, combine P and RK
DO KG = 1, KMAX+KADD
IA = ((KG-1+KN)*JSTRID + JN)*ISTRID
DO IG = 1+IN, IJSLB
II
= IA + IG
IM
= II - IL
UM
VM
WM

= 0.5*(RM(IM)+RM(II))
= 0.5*(RN(IM)+RN(II))
= 0.5*(RW(IM)+RW(II))

PERRO
RMCM
VFLOM

= 1.0/RO(II)
= AX(II)*UM + AY(II)*VM + AZ(II)*WM
= A(II)*(RMCM*PERRO-UROT(II))

PAM
FRM(II)
FRN(II)
FRW(II)
END DO
END DO

=
=
=
=

(P(II)+P23*RK(II))*A(II)
FRM(II) + PAM*AX(II) + VFLOM*UM
FRN(II) + PAM*AY(II) + VFLOM*VM
FRW(II) + PAM*AZ(II) + VFLOM*WM

IF (ISTRES .GT. 0) THEN ! 2-eq models with EARSM


DO KG = 1, KMAX+KADD
IA = ((KG-1+KN)*JSTRID + JN)*ISTRID
DO IG = 1+IN, IJSLB
II
= IA + IG
IM
= II - IL
C
C

Twice the mean values


(a_ij = 2.0*b_ij)
B11CM
=
B12CM
=
B13CM
=
B22CM
=
B23CM
=

of extra anisotrophy components b_ij


BIJ(II,1)
BIJ(II,2)
BIJ(II,3)
BIJ(II,4)
BIJ(II,5)

+
+
+
+
+

95

BIJ(IM,1)
BIJ(IM,2)
BIJ(IM,3)
BIJ(IM,4)
BIJ(IM,5)

Appendix D, Subroutine FLUXP

B33CM

=-BIJ(II,1) -BIJ(II,4) -BIJ(IM,1) -BIJ(IM,4)

The change of the momentum fluxes


RKA
= 0.5*(RK(II) + RK(IM))*A(II)
F2EX
= RKA*(AX(II)*B11CM+AY(II)*B12CM+AZ(II)*B13CM)
F3EX
= RKA*(AX(II)*B12CM+AY(II)*B22CM+AZ(II)*B23CM)
F4EX
= RKA*(AX(II)*B13CM+AY(II)*B23CM+AZ(II)*B33CM)

Updated momentum fluxes


FRM(II) = FRM(II) + F2EX
FRN(II) = FRN(II) + F3EX
FRW(II) = FRW(II) + F4EX
END DO
END DO
END IF
ELSE
! 0-eq models, use P alone
DO KG = 1, KMAX+KADD
IA = ((KG-1+KN)*JSTRID + JN)*ISTRID
DO IG = 1+IN, IJSLB
II
= IA + IG
IM
= II - IL
UM
VM
WM

= 0.5*(RM(IM)+RM(II))
= 0.5*(RN(IM)+RN(II))
= 0.5*(RW(IM)+RW(II))

PERRO
RMCM
VFLOM

= 1.0/RO(II)
= AX(II)*UM + AY(II)*VM + AZ(II)*WM
= A(II)*(RMCM*PERRO-UROT(II))

PAM
FRM(II)
FRN(II)
FRW(II)
END DO
END DO
END IF

=
=
=
=

P(II)*A(II)
FRM(II) + PAM*AX(II) + VFLOM*UM
FRN(II) + PAM*AY(II) + VFLOM*VM
FRW(II) + PAM*AZ(II) + VFLOM*WM

RETURN
END

96

Appendix E, Subroutine PEEKOO

SUBROUTINE PEEKOO(VIS,EPS2,PHITUR,DUIDXJ,BIJ,IMAX,JMAX,KMAX,
&
RK,PTUR)
*****************

C
C
C

A subroutine to calculate the production of turbulence with


the EARSM
2.11.2000 Ville H. / Laboratory of Aerodynamics
USE MAIN_ARRAYS , ONLY : MAXEB
USE CONSTANTS , ONLY : IN,JN,KN,EPS
IMPLICIT NONE
INTEGER , INTENT(IN) :: IMAX,JMAX,KMAX
INTEGER :: ISTRID,JSTRID,IL,IXCL,K,IA,IJ,L
REAL , INTENT(IN) , DIMENSION(MAXEB,9) :: DUIDXJ
REAL , INTENT(IN) , DIMENSION(MAXEB,5) :: BIJ
REAL , INTENT(IN) , DIMENSION(*) :: EPS2,VIS,RK,PHITUR
REAL , DIMENSION(*) :: PTUR
REAL :: P3,S11,S22,S33,S12,S13,S23
REAL :: RMUT,DIVVP3,HUU,HUV,HUW,HVV,HVW,HWW

One third
P3
= 1.0/3.0

Constants for the loops over the grid


ISTRID = IMAX + 2*IN
JSTRID = JMAX + 2*JN
IL
= ISTRID*JSTRID
IXCL
= JN*ISTRID + IN
DO K = 1,KMAX
IA = (KN+K-1)*IL
DO IJ = 1+IXCL,IL-IXCL
L = IA + IJ

Strain rate tensor components


S11
= DUIDXJ(L,1)
S22
= DUIDXJ(L,5)
S33
= DUIDXJ(L,9)
S12
S13
S23

= 0.5*(DUIDXJ(L,2) + DUIDXJ(L,4))
= 0.5*(DUIDXJ(L,3) + DUIDXJ(L,7))
= 0.5*(DUIDXJ(L,6) + DUIDXJ(L,8))

Effective eddy-viscosity coefficient


RMUT
= MAX(EPS2(L) - 1.0 , EPS)*VIS(L)

One third of the divergence of the velocity


DIVVP3 = P3*(S11 + S22 + S33)

Half of the Reynolds stresses (with rho) 0.5*rho*u_i*u_j


HUU
= MAX((RK(L)*(P3 + BIJ(L,1))
&
- RMUT*(S11 - DIVVP3)) , 0.0)
HVV
= MAX((RK(L)*(P3 + BIJ(L,4))
&
- RMUT*(S22 - DIVVP3)) , 0.0)
HWW
= MAX((RK(L)*(P3 - BIJ(L,1) - BIJ(L,4))
&
- RMUT*(S33 - DIVVP3)) , 0.0)
HUV
= RK(L)*BIJ(L,2) - RMUT*S12
HUW
= RK(L)*BIJ(L,3) - RMUT*S13
HVW
= RK(L)*BIJ(L,5) - RMUT*S23

Production of turbulence
PTUR(L) =-2.0*(HUU*S11 + HVV*S22 + HWW*S33
&
+ 2.0*(HUV*S12 + HUW*S13 + HVW*S23))

97

Appendix E, Subroutine PEEKOO

Forced laminarity by limited production if PHITUR is zero


PTUR(L) = PTUR(L)*PHITUR(L)
END DO
END DO
RETURN
END

98

Appendix F, Subroutine VELGRAD

SUBROUTINE VELGRAD(DUIDXJ,STRAIN,OHMI,OMX,OMY,OMZ,U,V,W,BLNK,
&
A1,A2,A3,VOL,A1X,A1Y,A1Z,A2X,A2Y,A2Z,A3X,A3Y,A3Z,
&
IMAX,JMAX,KMAX)
******************

C
C
C

A subroutine to calculate the velocity gradients and strain


and vorticity components.
18.11.2000 Ville H. / Laboratory of Aerodynamics
USE MAIN_ARRAYS , ONLY : MAXEB
USE CONSTANTS , ONLY : IN,JN,KN
IMPLICIT NONE
REAL
REAL
REAL
REAL
REAL

,
,
,
,
,

INTENT(IN) , DIMENSION(*) :: U,V,W,BLNK,A1,A2,A3,VOL


INTENT(IN) , DIMENSION(*) :: A1X,A1Y,A1Z,A2X,A2Y,A2Z
INTENT(IN) , DIMENSION(*) :: A3X,A3Y,A3Z
DIMENSION(*) :: STRAIN,OHMI,OMX,OMY,OMZ
DIMENSION(MAXEB,9) :: DUIDXJ

REAL :: U1M,V1M,W1M,U1P,V1P,W1P,S11,S12,S13,S22,S23,S33
REAL :: U2M,V2M,W2M,U2P,V2P,W2P,P3,PVOL,DIVVP3
REAL :: U3M,V3M,W3M,U3P,V3P,W3P
INTEGER , INTENT(IN) :: IMAX,JMAX,KMAX
INTEGER :: ISTRID,JSTRID,IL,IXCL,K,IA,IJ,L
C

One third
P3
= 1.0/3.0

Constants for the loops over the grid


ISTRID = IMAX + 2*IN
JSTRID = JMAX + 2*JN
IL
= ISTRID*JSTRID
IXCL
= JN*ISTRID + IN
DO K = 1,KMAX
IA = (KN+K-1)*IL
DO IJ = 1+IXCL,IL-IXCL
L
= IA + IJ

XI-direction
U1M = A1(L) *(U(L) + U(L-1))
V1M = A1(L) *(V(L) + V(L-1))
W1M = A1(L) *(W(L) + W(L-1))
U1P = A1(L+1)*(U(L) + U(L+1))
V1P = A1(L+1)*(V(L) + V(L+1))
W1P = A1(L+1)*(W(L) + W(L+1))
ETA-direction
U2M = A2(L)
*(U(L) + U(L-ISTRID))
V2M = A2(L)
*(V(L) + V(L-ISTRID))
W2M = A2(L)
*(W(L) + W(L-ISTRID))
U2P = A2(L+ISTRID)*(U(L) + U(L+ISTRID))
V2P = A2(L+ISTRID)*(V(L) + V(L+ISTRID))
W2P = A2(L+ISTRID)*(W(L) + W(L+ISTRID))
ZETA direction
U3M = A3(L)
*(U(L) + U(L-IL))
V3M = A3(L)
*(V(L) + V(L-IL))
W3M = A3(L)
*(W(L) + W(L-IL))
U3P = A3(L+IL)*(U(L) + U(L+IL))
V3P = A3(L+IL)*(V(L) + V(L+IL))
W3P = A3(L+IL)*(W(L) + W(L+IL))
PVOL
PVOL = 0.5*BLNK(IJ)/VOL(L)

99

Appendix F, Subroutine VELGRAD

Velocity gradients
DUIDXJ(L,1) = PVOL*(
&
A3X(L+IL)
*U3P
&
A2X(L+ISTRID)*U2P
&
A1X(L+1)
*U1P
DUIDXJ(L,2) = PVOL*(
&
A3Y(L+IL)
*U3P
&
A2Y(L+ISTRID)*U2P
&
A1Y(L+1)
*U1P
DUIDXJ(L,3) = PVOL*(
&
A3Z(L+IL)
*U3P
&
A2Z(L+ISTRID)*U2P
&
A1Z(L+1)
*U1P
DUIDXJ(L,4) = PVOL*(
&
A3X(L+IL)
*V3P
&
A2X(L+ISTRID)*V2P
&
A1X(L+1)
*V1P
DUIDXJ(L,5) = PVOL*(
&
A3Y(L+IL)
*V3P
&
A2Y(L+ISTRID)*V2P
&
A1Y(L+1)
*V1P
DUIDXJ(L,6) = PVOL*(
&
A3Z(L+IL)
*V3P
&
A2Z(L+ISTRID)*V2P
&
A1Z(L+1)
*V1P
DUIDXJ(L,7) = PVOL*(
&
A3X(L+IL)
*W3P
&
A2X(L+ISTRID)*W2P
&
A1X(L+1)
*W1P
DUIDXJ(L,8) = PVOL*(
&
A3Y(L+IL)
*W3P
&
A2Y(L+ISTRID)*W2P
&
A1Y(L+1)
*W1P
DUIDXJ(L,9) = PVOL*(
&
A3Z(L+IL)
*W3P
&
A2Z(L+ISTRID)*W2P
&
A1Z(L+1)
*W1P

- A3X(L)*U3M +
- A2X(L)*U2M +
- A1X(L)*U1M)
- A3Y(L)*U3M +
- A2Y(L)*U2M +
- A1Y(L)*U1M)
- A3Z(L)*U3M +
- A2Z(L)*U2M +
- A1Z(L)*U1M)
- A3X(L)*V3M +
- A2X(L)*V2M +
- A1X(L)*V1M)
- A3Y(L)*V3M +
- A2Y(L)*V2M +
- A1Y(L)*V1M)
- A3Z(L)*V3M +
- A2Z(L)*V2M +
- A1Z(L)*V1M)
- A3X(L)*W3M +
- A2X(L)*W2M +
- A1X(L)*W1M)
- A3Y(L)*W3M +
- A2Y(L)*W2M +
- A1Y(L)*W1M)
- A3Z(L)*W3M +
- A2Z(L)*W2M +
- A1Z(L)*W1M)

One third of the divergence of the velocity


DIVVP3
= P3*(DUIDXJ(L,1) + DUIDXJ(L,5) + DUIDXJ(L,9))

The strain components with the compressibility correction


S11
= DUIDXJ(L,1) - DIVVP3
S12
= 0.5*(DUIDXJ(L,2) + DUIDXJ(L,4))
S13
= 0.5*(DUIDXJ(L,3) + DUIDXJ(L,7))
S22
= DUIDXJ(L,5) - DIVVP3
S23
= 0.5*(DUIDXJ(L,6) + DUIDXJ(L,8))
S33
= DUIDXJ(L,9) - DIVVP3

The strain invariant sqrt(2*s_ij*s_ij):


STRAIN(L) = SQRT(2.0*(S11**2 + S22**2 + S33**2
&
+ 2.0*(S12**2 + S13**2 + S23**2)))
Vorticities
OMX(L)
OMY(L)
OMZ(L)
OHMI(L)

=
=
=
=

DUIDXJ(L,8) - DUIDXJ(L,6)
DUIDXJ(L,3) - DUIDXJ(L,7)
DUIDXJ(L,4) - DUIDXJ(L,2)
SQRT(OMX(L)**2 + OMY(L)**2 + OMZ(L)**2)

END DO
END DO
RETURN
END

100

HELSINKI UNIVERSITY OF TECHNOLOGY. LABORATORY OF AERODYNAMICS.


SERIES B

B-51

Juha Schweighofer and Antti Hellsten,


Computations of Viscous Flow around the HSVA-1 Tanker Using Two Versions of the k -
Turbulence Model, 1999.

B-50

Antti Hellsten,
On the Solid-Wall Boundary Condition of in the k - -Type Turbulence Models, 1998.

B-49

Antti Hellsten,
Implementation of a One-Equation Turbulence Model into the FINFLO Flow Solver, 1996.

B-48

Petri Kaurinkoski,
Development of an Equation of State for an Arbitrary Mixture of Thermally Perfect Gases to the
FINFLO Flow Solver, 1995.

B-47

Seppo Laine and Timo Sailaranta,


A Six Degree-of-Freedom Trajectory Model, 1995.

B-46

Harri Pitknen and Timo Siikonen,


Simulation of Viscous Flow in a Centrifugal Compressor, 1995.

B-45

Harri Pitknen and Seppo Laine,


Navier-Stokes Calculations of an Axisymmetric SOC-Projectile Using a Low-Reynolds-Number k-
Turbulence Model, 1995.

B-44

Jari Vilenius,
Katsaus ilmataistelun simulointiin, 1994

B-42

Reijo Lehtimki,
POISS3: a 3D Poisson Smoother of Structured Grids, 1993.

B-41

Jaakko Hoffren,
Computational Study of GA(W)-1 Airfoil Near Stall, 1993.

B-40

Huachen Pan,
Two-Dimensional Navier-Stokes Computations of Subsonic and Supersonic Flows through Turbine
Cascades, 1993.

B-39

Timo Sailaranta ja Seppo Laine,


Pervirtauksen vaikutus ammuksen ilmanvastukseen ja lentorataan, 1992.

ISBN 951-22-6824-8 (PDF)


ISSN 1456-6990

Vous aimerez peut-être aussi