You are on page 1of 109

MODELING TURBULENT FLOWS Introduction

7. MODELING TURBULENT FLOWS


7.1 Introduction
FIDAP provides extensive capabilities for the simulation of turbulent flows. This chapter
summarizes these capabilities and the steps required to set up and run a turbulent simula-
tion using FIDAP. Details on the theoretical background of both the equations solved and
the numerical techniques employed can be found in the FIDAP Theory Manual.
The purpose of this chapter is to help you make optimum use of the many FIDAP com-
mands and options available for obtaining convergent and realistic solutions to turbulent
flow problems. Some of the parameters that affect physical and numerical aspects of typi-
cal turbulent flow simulations are as follows:
· Boundary conditions
· Mesh density and spatial distribution
· Solution procedure
FIDAP requires only a few main commands and keywords for simulating turbulent flows.
The following table summarizes these commands.

Command Keyword Description

PROBLEM TURBULENT Indicates that the ensemble- or time-


averaged flow equations are to be solved.
Because these mean equations contain the
unknown turbulent fluxes, it is necessary
to select an appropriate turbulence model
by means of the VISCOSITY command
(see below).

VISCOSITY MIXING Specifies the use of a zero-equation type


turbulence model with the mixing length
computed by means of a user-subroutine,
USRMXL.

MIXLENGTH=v Specifies the use of a zero-equation type


turbulence model with automatic compu-
tation of the mixing length.

TWO-EQUATION Specifies the use of a two-equation type


turbulence model.

© Fluent Inc., Dec-98 7-1


Introduction MODELING TURBULENT FLOWS

Command Keyword Description

TURBOPTIONS STANDARD Specifies a particular two-equation type


EXTENDED turbulence model.
ANISOTROPIC
RNG
K-OMEGA

EDDYVISCOSITY BOUSSINESQ Specifies the eddy-viscosity constitutive


SPEZIALE relation to use in the turbulence model.
LAUNDER

This chapter presents information related to two different types of turbulent flow models:
· Zero-equation (Section 7.2)
· Two-equation (Section 7.3)
The bulk of the chapter is devoted to the use of the two-equation models. The reasons for
this focus are as follows:
· Two-equation models are more accurate and universal in their application and are
employed more frequently than are zero-equation models.
· There are considerably more physical and numerical issues involved in performing
turbulent flow simulations with a two-equation model, therefore it is more difficult
for the inexperienced user to obtain optimal solutions to such simulations.
Invoking a two-equation model entails the solution of two additional transport equations
which can significantly increase the CPU requirements of the numerical solution. In addi-
tion, the introduction of the k and e (or w ) equations significantly increases the nonline-
arity and coupling of the overall flow equations and, in general, destabilizes the con-
vergence characteristics of the numerical solution. By contrast, turbulent flow simulations
using zero-equation type models (where the value of the eddy viscosity is either fixed or
prescribed algebraically) exhibit relatively stable convergence characteristics similar to
those observed in the simulation of laminar flow problems.

7-2 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Introduction

To successfully use FIDAP to model turbulent flows, you must possess an adequate
knowledge of such flows and of turbulence models. Such knowledge will enable you to
choose the optimum turbulence model based on accuracy requirements and available CPU
resources. It will also allow you to correctly evaluate the solution to determine whether or
not the results are realistic. The task of verifying the solution is non-trivial and often
entails performing additional runs to check the sensitivity of the solution to boundary
conditions (such as the boundary conditions of k and e at the inlet plane to the computa-
tional domain) as well as other physical and/or numerical parameters. Sometimes it is
necessary to develop a “feel” for the numerics and the physics of turbulent flow simula-
tions—starting from simple flow problems and moving toward more difficult ones.

© Fluent Inc., Dec-98 7-3


Zero-Equation Flow Simulations MODELING TURBULENT FLOWS

7.2 Zero-Equation Flow Simulations


Zero-equation models employ only algebraic equations to describe the relationship
between the turbulent viscosity, m t , and the tangible flow quantities. Although there are
many different forms of zero-equation models, FIDAP employs a model known as the
mixing-length model. In the mixing-length model, m t is computed using an equation of
the form

c h
m t = rlm2 ui, j + u j ,i ui, j . (7.2.1)

(For a complete description of the significance and use of equation (7.2.1), see Chapter 10
of the FIDAP Theory Manual.)
Equation (7.2.1) involves a single unknown parameter known as the mixing length, lm ,
which can be thought of as the mean free path for the collision or mixing of globules of
turbulent fluid. The distribution of lm over the flow field must be prescribed with the aid
of empirical information.
The mixing-length model works well for relatively simple flows such as thin shear-layer
flows, wall boundary-layer flows, jets and wake flows, because lm can be specified by
simple empirical formulas in such situations. However, the model does not account for the
transport and history effects of turbulence. In particular, the model is not suitable when
processes of convective or diffusive transport of turbulence are important—such as in
rapidly developing flows, heat transfer across planes with zero velocity gradient, and
recirculating flows. More generally, the model is often difficult to apply in complex flows
because of the difficulties in specifying lm .

From a mathematical standpoint, a system of flow equations resulting from a zero-


equation type turbulent flow simulation is virtually identical to a system of laminar flow
equations with a constant or variable molecular viscosity. (Examples of laminar variable-
viscosity flows are flows that involve a temperature-dependent viscosity or non-Newto-
nian flows with shear-rate dependent viscosities.) Thus, as far as numerical aspects and
convergence characteristics are concerned, zero-equation type turbulent flow simulations
perform similarly to laminar flow simulations. In particular, if a well-defined set of
boundary conditions is employed, the numerical solution to a zero-equation type flow
model generally converges at a fast rate and produces a finely converged result in a rela-
tively small number of iterations. Additionally, the convergence characteristics of the
numerical solution are relatively insensitive to the following factors:
· Changes in the density and spatial distribution of the mesh
· Changes in the shape of the flow domain
· The starting guess solution (that is, the radius of convergence is relatively large)

7-4 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Zero-Equation Flow Simulations

The accuracy of the simulation, however, depends strongly on the adequacy of the zero-
equation turbulence model.

7.2.1 Mixing-Length Computation Methods

FIDAP provides two methods of specifying a zero-equation turbulence model:


· Automatic mixing-length computation
· User-subroutine mixing-length computation
The following sections describe each of these methods.

Automatic Mixing-Length Computation

When you specify the automatic mixing-length computation, FIDAP calculates the mixing
length at every point in the flow according to the following equation:
lm = min kln ,0.09lc (7.2.2)

where ln is the distance from the nearest wall, and lc is a characteristic length scale of the
flow. (For example, for an internal flow, lc is the half-width of the channel.) For any
given flow problem, FIDAP computes one global value of lc that is equal to the maximum
value of ln in the entire mesh. Although this global value is appropriate for problems
with simple geometries, it may not be representative for all flow regions in more compli-
cated geometries. For this reason, FIDAP provides the flexibility of overriding the value
of lc in selected flow regions.

To activate the automatic mixing-length computation, you must specify the keyword
MIXLENGTH=v on the VISCOSITY command, where the value v is the characteristic
length scale, lc .
· If you specify a MIXLENGTH value of zero, or do not specify a value, FIDAP
employs the computed global value of lc .
· If you specify a non-zero value for v, FIDAP overrides the computed global value
of lc .

If you define VISCOSITY model sets for two or more fluids, the MIXLENGTH keyword
allows you to explicitly specify the characteristic length scale lc for the different FLUID
entities in various regions in a geometrically complex flow.

© Fluent Inc., Dec-98 7-5


Zero-Equation Flow Simulations MODELING TURBULENT FLOWS

User-Subroutine Mixing-Length Computation

In lieu of the automatic computation of mixing length, FIDAP allows you to explicitly
specify the mixing-length computation by means of the user subroutine, USRMXL. To
employ a user-subroutine mixing-length computation, you must specify the MIXING
keyword on the VISCOSITY command and provide the USRMXL subroutine. For a
complete description of the USRMXL subroutine and its use, see Chapter 9 of the
FIPREP User’s Manual. (NOTE: A large number of zero-equation type turbulence
models are available in the literature.)

7.2.2 Specification Procedures

The steps required to specify a zero-equation turbulence model are as follows:

Step Description

1 Specify a TURBULENT analysis on the PROBLEM command.

2 · For an automatic mixing-length computation, specify the MIXLENGTH


keyword on the VISCOSITY command.
· For a user-subroutine mixing-length computation, specify the MIXING
keyword on the VISCOSITY command and provide a subroutine
(USRMXL) to compute the mixing length.

3 (MIXLENGTH option only)


Specify WALL boundary entities on any portions of the boundary which are
not inflow, outflow or symmetry boundaries. (NOTE: FIDAP automatically
assumes that the boundary of any SOLID or POROUS entity is a wall for
the purposes of the mixing-length computations, therefore WALL entities
are not required on these boundaries.)

7-6 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Two-Equation Flow Simulations

7.3 Two-Equation Flow Simulations


Two-equation turbulence models employ two extra partial differential equations to
describe the relationship between the turbulent viscosity, m t , and the tangible flow quan-
tities. FIDAP provides five two-equation turbulence models and three eddy-viscosity
constitutive relations (see the following table).

Turbulence Models Eddy-Viscosity Constitutive Relations

· Standard k - e (default) · Boussinesq (Default)


· Extended k - e · Speziale
· RNG k - e · Launder
· Anisotropic k - e
· Wilcox Low-Re k - w

For most cases that involve high-Re flows, the model defaults (standard k - e model and
Boussinesq constitutive relation) produce satisfactory results. For low-Re flows, the rec-
ommended practice includes the use of the k - w model and Boussinesq relation. For a
complete description of the models and relations listed above and their use in turbulence
modeling, see Chapter 10 of the FIDAP Theory Manual.
Employing a two-equation turbulence model significantly increases the required CPU
resources for the solution of a given flow problem compared to the corresponding simula-
tion using a zero-equation type model. Moreover, because of the significantly stronger
nonlinear and intercoupled nature of the system of flow equations, it becomes somewhat
more difficult to obtain a fully converged solution. In general, the convergence character-
istics of two-equation simulations are less stable than are those of zero-equation simula-
tions and are somewhat sensitive to the numerical and physical parameters involved in the
numerical solution. In the following subsections, these various parameters are highlighted
and some guidelines are provided as to how to best manipulate them (by means of the
available FIDAP commands) to obtain a stable numerical solution.

© Fluent Inc., Dec-98 7-7


Two-Equation Flow Simulations MODELING TURBULENT FLOWS

7.3.1 Boundary Conditions

One important difference between two-equation simulations and zero-equation simula-


tions is that, for a two-equation simulation, appropriate boundary conditions must be pre-
scribed on the boundaries of the computational domain for the primary turbulence
variables—that is, k and e (or w ). In general, five types of computational boundaries are
encountered in typical flow simulations (see Figure 7-1). They are:
· Inlet
· Outlet
· Symmetry
· Wall
· Entrainment

Figure 7-1: Typical computational flow domain boundaries

The following subsections describe the boundary conditions that must be applied for the
primary turbulence variables at the boundaries listed above.
NOTE: To specify a boundary condition for the turbulent frequency ( w ) for the k - w
model, you must input the corresponding value for the dissipation, e ( = kw ) by means of
the DISSIPATION keyword on the BCNODE command. When you specify the use of the
k - w model, FIDAP automatically converts the input e value to its corresponding turbu-
lent frequency (w = e k).

7-8 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Two-Equation Flow Simulations

Inlet boundaries

Inlet boundaries are typically positioned upstream of the regions of interest and are
located in areas where the flow field is unperturbed by any nearby obstacles. Dirichlet
(that is, prescribed or essential) boundary conditions for k and e (or w ) must be em-
ployed on inlet boundaries.
Strictly speaking, the levels and the shapes of the profiles of k and e (or w ) at the inlet
boundary are unique for every flow problem. Ideally, they should be obtained from
experimental measurements, but such experimental data is rarely available for typical
simulations. Moreover, a precise set of laws does not exist from which appropriate
profiles for k and e (or w ) can be derived for all possible flow scenarios.
The two most commonly occurring situations are as follows:
· External/unconfined flows
· Internal or partially confined flows
The following paragraphs describe the procedures required to obtain reasonable estimates
for the characteristic k and e scales for each of the flow situations listed above.

External/Unconfined Flows

Estimating k
To obtain a reasonable estimate of k for external or unconfined flows, you must first
specify a characteristic velocity scale of the mean flow—for example, the free stream
velocity, u¥ . Then you can obtain the characteristic value of k from

k = au¥2
(7.3.1)

where a is a turbulence intensity which assumes the following range of values

{
- O(0.1)
a~
; free - shear flows (wakes, jets, etc.)
O(0 ~ 0.001) ; shear - free flows

In some shear-free situations—as is often the case in wind tunnel flows—an alternative
definition of turbulence intensity, I, is used. In such cases, the turbulence intensity is
defined as

eu ¢ j
1
2
2

I= (7.3.2)

and a value for k can readily be arrived at from the following expression,

© Fluent Inc., Dec-98 7-9


Two-Equation Flow Simulations MODELING TURBULENT FLOWS

b g
k = 1.5 Iu¥
2
. (7.3.3)

Estimating e
You can obtain the characteristic value of e in one of two ways, depending on whether a
characteristic eddy length scale d l can be identified at the inlet boundary (free-shear flow)
or not (shear-free flows).
If the eddy length scale can be identified, e can be obtained from the expression
3 3
k2~ k2
e= - (7.3.4)
d l 0.1d

where d is the characteristic width of the shear layer at the inlet plane (see Figure 7-2).
If the eddy length scale cannot be identified—that is, a characteristic length scale of the
mean flow or turbulence fields is not available—a characteristic value of e can be
obtained from the expression

k2
e = rcm (7.3.5)
Rm m

where Rm = m t m is the ratio between the turbulent and laminar viscosities. In general,
Rm varies considerably from one flow problem to another. However, for most problems a
“reasonable” value of Rm can be arrived at after some trial and error. For example, for
af
free-shear flows with very low levels of turbulence Rm ~ O 1 . For most typical flows Rm
~ O(10 ) – O(10 ).
1 2

Specifying k and e
The characteristic values of k and e obtained from the procedures listed above can be
specified as Dirichlet boundary conditions at the inlet plane by means of the KINETIC
and DISSIPATION keywords, respectively, on the BCNODE command. Moreover, if
equations (7.3.1) and (7.3.4) are used to arrive at characteristic inlet k and e values, they
can be applied simultaneously together with the inlet velocity boundary condition using
the INTENSITY and LENGTH keywords of the BCNODE command. The value of the
INTENSITY keyword on the BCNODE command is the ratio of k to the square of the
characteristic velocity scale, u¥ , expressed as a percentage. Thus, with respect to equation
(7.3.1), the value of the INTENSITY keyword equals a ´ 100. The value of the LENGTH
keyword is the size of the characteristic eddy length scale, d l , that appears in equation
(7.3.4).

7-10 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Two-Equation Flow Simulations

δ
δ

δ
δ

Figure 7-2: Schematic definition of d , the shear layer width, for various types of flows

© Fluent Inc., Dec-98 7-11


Two-Equation Flow Simulations MODELING TURBULENT FLOWS

Internal or Partially Confined Flows

Estimating k and e
In the case of internal (fully confined) or partially confined flows (for example, boundary
layer flows), the following expressions can be used to arrive at appropriate profiles for k
and e :

F
k = c Gl
du IJ 2

H K
1

m
2
m (7.3.6)
dy

F
e = c k Gl
du I
-1

H dy JK
m
2 2
m × (7.3.7)

where lm is a mixing-length expression appropriate to the current model, u is the stream-


wise velocity component at the inlet plane, and y is the normal coordinate axis to the
nearest wall (see Figure 7-3).
If the inlet plane is located in a region that is remote from flow obstructions, the stream-
wise velocity profile at the inlet plane can be closely approximated by a power-law profile
of the form

u y FG IJ 1
h

ur
=
d H K , (7.3.8)

where the exponent h is Reynolds number dependent and is obtained from the following
table.

rurd m h
4 ´ 103 6.0
2.3 ´ 104 6.6
1.1 ´ 105 7.0
1.1 ´ 106 8.8
10
2 ´ 106
10
3.2 ´ 106

7-12 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Two-Equation Flow Simulations

y u

ur

ur

u
y δ

Figure 7-3: Flow configuration at inlet boundary for fully- and semi-confined flows

Specifying k, e , and u
The above profiles of k, e and u can be imposed as Dirichlet boundary conditions on the
inlet boundary in FIDAP using the user-subroutine option of the BCNODE command.
For a description of the BCNODE command and related user subroutines, see Chapter 7
of the FIPREP User’s Manual.

© Fluent Inc., Dec-98 7-13


Two-Equation Flow Simulations MODELING TURBULENT FLOWS

Outlet and Symmetry Boundaries

The Neumann (that is, the zero-gradient or zero-flux) boundary condition is the most
appropriate for k and e on outlet and symmetry boundaries. As part of the finite element
discretization, FIDAP automatically assigns Neumann boundary conditions on portions of
the computational boundaries to those flow variables that have not been explicitly given
Dirichlet or gradient boundary conditions by means of the BCNODE and BCFLUX
commands, respectively. Thus, in the context of FIDAP, Neumann boundary conditions
are imposed on k and e on a given outlet or symmetry plane if no explicit boundary con-
ditions for k and e are applied at that location.
CAUTION: In designing the computational mesh, you must ensure that the outflow
boundary is placed at a downstream location that is sufficiently far from regions of the
flow where large perturbations occur in the flow field. For example, if flow over a back-
ward facing step is being simulated, the flow exit plane of the computational domain
should be placed sufficiently downstream of the point of reattachment to avoid any possi-
ble interaction between this boundary and the flow patterns in the recirculation zone.
Specifically, in the case of turbulent flow over a backward facing step, the stream-wise
extent of the recirculating bubble (that is, the reattachment length) is approximately seven
step heights. The exit plane must be placed at least one and a half recirculating lengths
(that is, about 11 step heights) downstream of the reattachment point.

Wall Boundaries

In turbulent flows that are bounded by walls, the near-wall modeling methodology must
be applied along those portions of the computational boundary that coincide with the solid
walls. This specialized methodology is completely transparent to the FIDAP user. (For a
complete description of the near-wall modeling methodology, see Chapter 10 of the
FIDAP Theory Manual.)
To apply the near-wall model, you must assign a WALL boundary entity along the portion
of the computational boundary that coincides with the solid boundary. When a WALL
entity is assigned, FIDAP searches all FLUID entities and identifies, as special near-wall
elements, those FLUID elements the sides of which coincide with the WALL boundary
elements. You must also apply the appropriate boundary conditions for those mean flow
equations (for example, momentum, temperature, and/or species) that are being solved as
part of the flow model.

7-14 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Two-Equation Flow Simulations

Momentum Boundary Conditions

To specify a problem that involves an isothermal flow with a stationary solid wall, you
must assign zero values for velocity components at the wall. If moving walls are present
and/or if there is transpiration at the wall, the appropriate non-zero velocity boundary
conditions must be imposed at the wall. (NOTE: If you prescribe non-zero velocity
boundary conditions along WALL boundaries, the velocity components on any node lying
on these boundaries should be entered in terms of the global coordinate system, rather
than a local normal-tangential system such as is used for SLIP boundaries.)

Temperature and Species Boundary Conditions

If heat or mass transfer is present in the problem, an appropriate boundary condition for
the temperature or species equation is needed. The boundary condition can take the form
of a prescribed temperature (or species concentration) or a prescribed heat (or mass) flux.
In the case of a conjugate heat and/or mass transfer problem, where the boundary between
the fluid and solid regions is internal to the computational domain, no explicit boundary
conditions need to be applied at the fluid/solid interface for the energy or species equa-
tions, because the values of temperature and species concentration at this interface are
computed as part of the numerical solution.
NOTE: When WALL boundary entities are employed in a model, FIDAP uses a boundary
unit normal vector to compute the characteristic cross-flow widths of the special near-wall
elements. On external boundaries of the computational domain, the direction of the unit
normal vector is uniquely defined as pointing away from the computational domain. On
internal boundaries, however, the direction of this normal vector is not uniquely defined,
and you must provide a direction. For internal boundaries involving WALL entities, you
should orient the unit normal vector such that it is pointing away from the FLUID entity
and toward the SOLID entity. This orientation is achieved by means of the ATTACH
keyword when the WALL boundary entity is defined using the ENTITY command. The
value of the ATTACH keyword must be set to the name of the FLUID entity. For a
detailed description of the ATTACH keyword, see Chapter 6 of the FIPREP User’s
Manual.

Porous Media Boundary Conditions

Internal interface boundary conditions for turbulent flows that involve porous media (with
or without heat/mass transfer) are similar to those described above. However, an fluid-
porous interfaces differ from fluid-solid interfaces in that they involve flow across the
interface. Thus, no velocity boundary condition should be explicitly applied at such an
interface, because the velocity at that location is computed as part of the numerical
solution. If significant shearing of the flow is anticipated at the fluid-porous interface, it is
recommended that WALL boundary entities be employed along the interface.

© Fluent Inc., Dec-98 7-15


Two-Equation Flow Simulations MODELING TURBULENT FLOWS

Boundary Conditions on k and e

It is unnecessary to apply k and e boundary conditions on the “wall” portion of the com-
putational domain, because this is done automatically within FIDAP as part of the near-
wall model. The k and e boundary conditions that FIDAP applies are described in
Chapter 10 of the FIDAP Theory Manual.

Entrainment Boundaries

Entrainment boundaries typically occur in external or semi-confined flow configurations.


There are two common types of entrainment boundaries, each of which can be distin-
guished from the other by the size of the ratio of the entrainment velocity to the tangential
velocity. The two types are as follows:
· Small ratio (see Figure 7-4)
· Large ratio (see Figure 7-5)

Small-Ratio Entrainment Boundary

Figure 7-4 shows two examples of boundaries that are characterized as having a small
entrainment velocity component (that is, the velocity component normal to the boundary
plane) relative to the tangential velocity component. The figure shows that entrainment on
such boundaries can be both in and out of the flow domain.
Zero-traction boundary conditions are appropriate for the components of the momentum
equation. The appropriate boundary conditions for k and e on these boundaries are
Neumann (that is, zero-gradient) conditions. In some cases, “harder” Dirichlet boundary
conditions for k and e can be applied on these entrainment planes (the characteristic
values of k and e employed at the inlet plane can be used for this purpose), but these are
not as appropriate as the Neumann boundary conditions.

7-16 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Two-Equation Flow Simulations

Figure 7-4: Entrainment boundaries on which entrainment velocity is small


compared to tangential velocity

© Fluent Inc., Dec-98 7-17


Two-Equation Flow Simulations MODELING TURBULENT FLOWS

Figure 7-5: Entrainment boundary on which entrainment velocity is large


compared to tangential velocity

Large-Ratio Entrainment Boundary

Figure 7-5 shows an entrainment boundary that is characterized as having a large entrain-
ment velocity that points into the flow domain and is considerably larger than the tangen-
tial velocity component. The appropriate boundary conditions for the velocity components
are a zero tangential velocity component and a zero normal traction condition for the
normal velocity component. The appropriate boundary conditions for k and e are
Dirichlet conditions characterizing the state of turbulence in the free stream regions
outside the flow domain. Equations (7.3.1) to (7.3.5) can be used to arrive at the charac-
teristic free stream values of k and e . If the free stream is totally turbulence free (that is,
if it is laminar), k and e can be set to zero on the entrainment boundary.
7.3.2 Initial Conditions for k and e

Experience has shown that the convergence characteristics of most k - e runs are consid-
erably improved when non-zero initial guess fields are used for k and e . The recom-
mended approach is therefore to impose constant non-zero initial fields of k and e via the
KINETIC and DISSIPATION keywords on the ICNODE command, irrespective of
whether a steady-state or transient simulation is being performed. The initial levels of k
and e may be set equal to the characteristic values that are obtained from equations
(7.3.1) to (7.3.5). On the other hand, if equations (7.3.6) and (7.3.7) are being used for
prescribing profiles of k and e at the inlet boundary, intermediate values from these pro-
files can be used for the initial guess values.

7-18 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Two-Equation Flow Simulations

7.3.3 Solution Algorithms and Strategies

The type of solution approach (or strategy or algorithm) appropriate for any particular
turbulent flow problem depends on the following factors:
· Physical attributes of the flow (for example, steady or transient, 2-D or 3-D)
· Availability of computational resources (for example, memory limits may prohibit
you from using a faster but more memory intensive solution strategy)
The following sections describe FIDAP solution algorithms and solution strategies for
turbulent flows. For a complete description of the solution algorithms available in FIDAP,
see Chapter 7 of the FIDAP Theory Manual.

Solution Algorithms

FIDAP provides two classes of algorithms for the numerical solution of the typical set of
discretized equations which result from the application of the Galerkin finite element
method to the governing flow equations.
· Implicit algorithms
· Explicit algorithm

Implicit Algorithms

The set of discretized algebraic equations that results from the stationary form of the flow
equations (or from the transient form of these equations when an implicit time integration
scheme is employed) is implicit for all flow variables. This means that non-trivial matrix
systems must be inverted to compute all the nodal unknown flow variables, such as
velocities, pressure, and temperatures. These matrix systems should be solved with one of
the implicit algorithms available in FIDAP.
The implicit algorithms are used for solving most flow problems—whether they are tur-
bulent or not. This is because most flows considered are either steady, or in the case of
transient flows, are integrated implicitly in time—primarily to avoid stability limits on the
size of the time step, and/or to use the variable time stepping feature which is only avail-
able with the implicit time integration schemes.
For the implicit algorithms, FIDAP provides two solution approaches for solving the
nonlinear equations that result from the finite element discretization:
· Fully coupled approach
· Segregated approach

© Fluent Inc., Dec-98 7-19


Two-Equation Flow Simulations MODELING TURBULENT FLOWS

Fully Coupled Approach


In the fully coupled approach, discretized equations are assembled into one large global
matrix equation system for later simultaneous solution. For the fully coupled approaches,
FIDAP allows the user some flexibility in designing the structure of the solution algo-
rithm. One may adopt either a mixed or a penalty formulation for the pressure-velocity
coupling. For nonlinear problems, two basic choices of linearization schemes are
available:
· Fixed-point successive substitution scheme
· One of the Newton-type schemes (that is, Newton-Raphson, Modified-Newton or
Quasi-Newton)
The global system of linear equations that results at each iteration from the fully coupled
approach is solved using direct LU factorization.
The fully coupled approach is the most direct solution approach used in FIDAP. Its great-
est advantage is that it typically takes the least number of iterations to arrive at a solution.
(The solution to linear flow problems is obtained in one iteration.) Moreover, this
approach is usually very efficient in terms of CPU time for most two-dimensional flows.
For very large, two-dimensional flows and most three-dimensional flows, however, the
CPU time requirements may become excessive. The main disadvantage of the fully
coupled approach is that its memory and disk storage requirements are the largest of all
solution approaches available in FIDAP.
Segregated Approach
In the segregated approach, the global matrix system is never directly constructed.
Instead, the discretized equations associated with each primary flow variable are assem-
bled in smaller matrix equation systems.
Unlike the fully coupled approach, the structure of the segregated approach is essentially
fixed and is not subject to user control. Thus, when this approach is selected, FIDAP
automatically adopts the mixed velocity-pressure formulation. Moreover, if the system of
flow equations is nonlinear, the fixed-point (or successive substitution) scheme is auto-
matically used to linearize the various nonlinear sub-matrix systems.
The default method employed in FIDAP for solving the various linear equation systems
resulting from the segregated solution approach is the direct LU factorization method. In
order to offset the high costs of the direct solver in large-scale 3-D problems, FIDAP also
provides the option of solving the above linear equation systems using iterative methods.
These solvers together with various preconditioning strategies are described in Chapter 7
of the FIDAP Theory Manual.

7-20 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Two-Equation Flow Simulations

The greatest advantage of the segregated approach is that it substantially reduces memory
and disk storage requirements compared to the fully coupled approach. This relative
resource requirement becomes more pronounced (in favor of the segregated approach) as
the number of transport equations in the flow model is increased. (This is especially true
when the iterative linear equation solvers are being employed.) However, because of its
decoupled nature, the segregated approach tends to require more iterations to produce a
solution than does the fully coupled approach. This is also true for linear flow problems,
where typically more than one iteration is required to obtain a solution. Our experience
indicates that, from the standpoint of CPU time, the segregated approach is not as effi-
cient as the fully coupled approach for most two-dimensional problems involving few
transport equations. However, the relative efficiency of the segregated approach improves
dramatically with an increase in the number of transport equations in the flow model. A
similar dramatic improvement is also gained when the dimensionality of the problem
increases from two to three. For example, for three-dimensional k- e simulations, the seg-
regated approach not only requires much less memory but also consumes substantially
less CPU time.

Explicit Algorithm

For the explicit algorithm, the set of discretized equations that results from the transient
form of the flow equations when the forward Euler time integration scheme is employed
for the temporal terms is explicit for all flow variables except the pressure. This means
that a non-trivial matrix system is encountered only for the pressure unknowns. The
matrix systems resulting for all the other flow variables are diagonal and are trivial to
invert. These matrix systems are solved as part of the explicit algorithm in FIDAP.
The explicit algorithm is automatically invoked in FIDAP when the forward Euler scheme
is used to integrate the transient flow equations in time. In contrast to the implicit algo-
rithms, which can be used for either the steady or transient flow equations, the explicit
algorithm is essentially a transient algorithm and is only appropriate for solving the time-
dependent version of the flow equations. Also, as is the case with the segregated algo-
rithm, the structure of the explicit algorithm is fixed within FIDAP and is not subject to
user control. This algorithm automatically uses a mixed velocity-pressure formulation,
and only requires the solution of one non-trivial matrix problem for the pressure
unknowns. This system of linear equations for the pressure unknowns is solved using
direct LU factorization. (Iterative solvers may not be used here.)
An advantage of this algorithm is that it has a relatively small memory requirement. Its
main disadvantage is that the size of the time step is limited by various stringent stability
criteria. Thus, with this algorithm you typically need to use a relatively large number of
time steps to cover a time interval which is significant with respect to the characteristic
time scale of the mean flow process.

© Fluent Inc., Dec-98 7-21


Two-Equation Flow Simulations MODELING TURBULENT FLOWS

Solution Strategies

The solution strategy is probably the most important factor in two-equation simulations of
turbulent flows. FIDAP solution strategies can be classified as belonging to one of two
general types.
· Most 2-D flows
· Large 2-D and most 3-D flows

Most 2-D Flows

For most 2-D flows in which the global system matrix can be accommodated in computer
memory and peripheral disk storage, the fully coupled implicit algorithm should be used,
because it tends to be the most efficient for such problems. Also, the penalty formulation
should usually be used with this approach, because it removes the pressure degree of free-
dom from the global system matrix. (In the penalty formulation, pressure is recovered in a
post-processing phase of the computation.) Moreover, the successive substitution scheme
should be used for linearizing the system matrix. The Newton-type schemes appear to be
inappropriate for the highly nonlinear turbulent flow equations resulting from the k - e
model and, for this reason, are disallowed for turbulent flows.
To invoke the fully coupled implicit algorithm in conjunction with the successive substi-
tution scheme, you must specify the S.S. keyword on the SOLUTION command.

Large 2-D and Most 3-D Flows

For large 2-D flow problems and most 3-D problems, the segregated algorithm should be
used. It is invoked using the SEGREGATED keyword on the SOLUTION command. If
iterative linear equation solvers are to be employed, the recommended choice of solvers
and preconditioners is as follows.
· Use the CR method for the symmetric pressure-type equation and the CGS method
for the non-symmetric advection-diffusion equations resulting from the various
conservation equations.
· In conjunction with these solvers, use the SSOR preconditioners with the CR
method and the diagonal preconditioners with the CGS method.
· Use relatively tight tolerances (for example, 10-6) to ensure good overall conver-
gence behavior and solution quality.
The resulting SOLUTION command has the following keyword configuration:
SOLUTION(SEGR=n, CR=n, CGS=n, NCGC=1E-6, SGCG=1E-6, PRECON=21)

7-22 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Two-Equation Flow Simulations

Implicit Algorithms
When using the implicit algorithms, it is very important to specify the relaxation of the
numerical solution in the iterative solution of the nonlinear turbulent flow equations.
Relaxation can be introduced in one of two ways:
· The ACCF keyword of the SOLUTION command
· The RELAXATION command
When the ACCF keyword is used, all flow equations present in the model are relaxed
equally and the relaxation factor which is used is equal to the value of the ACCF
keyword. With the fully coupled algorithm, this value may range from its default value of
zero for very simple unidirectional and confined flows to about 0.9 for highly complex
recirculating flows with strong intercoupling effects due to buoyancy or swirl or both.
However, a more typical range is between 0.4 and 0.7.
The RELAXATION command can be used to selectively relax the various flow
equations. This is the recommended approach for the segregated algorithm. The relaxation
factors typically used for this algorithm are about 0.1 to 0.4 for velocities, 0 to 0.2 for
pressure, 0 to 0.1 for temperature and species concentration, and 0.1 to 0.3 for k and e .
If the segregated algorithm is used and no relaxing is explicitly specified by means of the
ACCF keyword or the RELAXATION command, FIDAP automatically employs a set of
non-zero default relaxation factors for the various flow equations that are present in the
flow problem. (For more information, see the FIPREP User’s Manual). Note that the
ACCF keyword takes precedence over the RELAXATION command. If this keyword is
present on the SOLUTION command and has been assigned a value, this value is the
relaxation factor that will be used for all flow equations. Any values entered using the
RELAXATION command are ignored.

Explicit Algorithm
The explicit algorithm should be used for those transient flow problems where very small
time steps are required to accurately resolve the temporal character of the flow. For prob-
lems of moderate size, the explicit algorithm may require the least CPU time per time
step. It should be noted, however, that cases requiring the use of very small time steps
seldom occur in practice. As a result, the explicit algorithm is rarely needed for turbulent-
flow computations.

© Fluent Inc., Dec-98 7-23


Two-Equation Flow Simulations MODELING TURBULENT FLOWS

7.3.4 Mesh Density and Distribution

In general, a denser computational grid should be employed for a two-equation simulation


compared to the mesh of the corresponding laminar or zero-equation simulation. This is
due to the fact that, in a typical flow problem, the turbulence variables undergo much
sharper spatial variations and involve considerably more detailed features than do the
mean flow variables (that is, the velocity and temperature fields). Thus, if a grid-inde-
pendent solution to the mean flow field is to be obtained, the computational mesh must be
fine enough to resolve the details of the k and e (or w ) fields. Care must also be exer-
cised in the design of the computational mesh to ensure a smooth spatial distribution of
nodal points throughout the flow domain. Abrupt jumps in mesh density may lead to
spurious spatial oscillations in the flow variables, especially if they occur along the direc-
tion of the flow in regions where the grid Reynolds numbers are large. In extreme cases,
these so-called “wiggles” cause the numerical solution to diverge.

7.3.5 Solution Stability

Sources of Instability

There are three main sources of instability that, if untreated, adversely affect typical two-
equation simulations.
· Instability associated with the dissipation (or sink) terms in the k and e equations
· Instability associated with the advection terms in the k and e equations.
· Instability that ensues when the k- e model is used in the prediction of flows con-
taining both turbulent and laminar regions
These instabilities manifest themselves in terms of non-physical negative k and e (or w )
values (k, e , and w are physically positive definite quantities which can never assume
negative values in the real world) and/or highly unrealistic turbulence time and length
scales resulting from very small k and e values in flow regions where turbulence levels
have practically collapsed. The seat of these instabilities lies in the k and e (or w )
equations.

Instability Due to Dissipation Terms

The first source of instability is that associated with the dissipation (or sink) terms in the
k and e (or w ) equations. Physically, these terms act to maintain finite levels of k and e
(or w ). In the absence of these terms k and e levels grow uncontrollably (and exponen-
tially) due to the generation (or source) terms of the k and e (or w ) equations.

7-24 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Two-Equation Flow Simulations

During the course of a typical numerical solution, while the interim solution field is sig-
nificantly different from the fully converged solution, the dissipation terms may strongly
dominate the generation terms and can momentarily produce destabilizing negative nodal
values of k and e (or w ). These negative values are destabilizing, because they change
the polarity of crucially important processes in the k and e (or w ) equations. For
example, the turbulent viscosity will become negative causing highly destabilizing nega-
tive diffusion. The source and sink terms will change polarity—for example, the turbulent
generation terms, instead of extracting turbulence energy from the mean flow process,
will extract energy from the turbulence field and impart it to the mean flow. Similarly the
sink terms, instead of dissipating turbulence energy, will generate turbulence energy (a
physical impossibility).

Instability Due to Advection Terms

The second source of instability is associated with the advection terms in the k and e (or
w ) equations. It is well known that, at large Reynolds numbers, these terms produce
stream-wise oscillations in the corresponding flow variables if they are approximated
using accurate non-diffusive discretization operators. (Such operators automatically result
from the application of the Galerkin finite element method to the flow equations.) Nega-
tive nodal values of k and e (or w ) could therefore result if these oscillations are large
compared to the local values of k and e (or w ).

Instability in Flows with Both Laminar and Turbulent Regions

The third source of instability ensues when a two-equation model is used in the prediction
of flows containing both turbulent and laminar regions. A typical example of this is in ex-
ternal aerodynamic problems where the free-stream flow is turbulence-free and the flow
surrounding the body is fully turbulent. The k and e equations of the standard high-
Reynolds-number k - e turbulence model become anomalous in the turbulence-free
regions. These equations contain terms involving ratios between k and e (that is, k2 e ,
k e , e k and e 2 k) which are clearly indeterminate in laminar flow regions where turbu-
lence is not present (that is, k and e are zero). Numerically, these ratios become very sen-
sitive to noise level variations in k and e and begin to oscillate violently (and
anomalously) from one nodal point to another. This can have a devastating effect on the
numerical stability of the computation, as this unstable behavior quickly spreads to the
fully turbulent flow regions and in a matter of a few iterations totally contaminates the nu-
merical solution.

© Fluent Inc., Dec-98 7-25


Two-Equation Flow Simulations MODELING TURBULENT FLOWS

Stabilization Techniques

Two basic techniques are available in FIDAP for suppressing the instabilities described
above.
· Upwinding
· Clipping

Upwinding

The purpose of upwinding is to stabilize advection terms in flow regions of high


advection. FIDAP provides three types of upwinding:
· Streamline
· First-order
· Hybrid
Streamline upwinding stabilizes the high-order symmetric advection operators that arise
in the Galerkin method by explicitly adding numerical diffusion only along the flow
direction. In doing so, it helps to suppress the stream-wise oscillations in the various flow
variables (including k and e ) that occur in advection dominated flow regions. In first-
order upwinding, the convection operator is decomposed in the principal directions, then
full upwind differencing is applied in the respective principal directions. Hybrid upwind-
ing is an extension of first-order upwinding in which a blending parameter (a ) that is
used in the first-order scheme is dynamically computed during the course of the computa-
tions rather than being fixed by the user. (For detailed descriptions of the three upwinding
schemes, see Chapter 7 of the FIDAP Theory Manual.)
Streamline upwinding is the recommended upwinding technique for most FIDAP
problems. When you specify a problem as TURBULENT, FIDAP activates the streamline
upwinding scheme and employs default upwind factors of unity (1).
There are two ways to manually invoke streamline upwinding in FIDAP.
· The UPWINDING keyword on the OPTIONS command
· The UPWINDING keyword on the OPTIONS command in conjunction with the
UPWINDING command
When the UPWINDING keyword on the OPTIONS command is used alone, all flow
equations present in the flow model are upwinded by equal amounts, and the upwinding
factor used is equal to the value of the UPWINDING keyword. A typical value for the
upwinding factor for turbulent flows is unity for k - e simulations, because they are
usually advection dominated. Lower values may be used if advection effects are not

7-26 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Two-Equation Flow Simulations

strong. In extreme cases, values larger than one could be used. However, this value should
not exceed 2.
The UPWINDING command in addition to an OPTIONS(UPWINDING) command with
no keyword value may be used to selectively upwind the various equations present in the
flow model. The upwinding factors typically used with this command for turbulent flows
are about unity for the mean flow variables (velocities, temperature and species concen-
tration) and between one to five for k and e . Note that an UPWINDING keyword value
on the OPTIONS command takes precedence over the UPWINDING command. If this
keyword value is present on the OPTIONS command, the upwinding factors assigned via
this keyword will be used for all flow equations and any such values entered using the
UPWINDING command are ignored.

Clipping

Clipping is a procedure that avoids the first and third types of instability described above
by ensuring that nodal values of k and e do not fall below pre-assigned lower bound
positive values. This procedure is described under the VISCOSITY command in the
FIPREP User’s Manual.
The lower bound values below which nodal values of k and e are clipped are set by
default to be ten million times smaller than the maximum nodal values of k and e . This is
adequate for virtually all cases. However, in flows involving a combination of exception-
ally large mean flow Reynolds numbers (that is, Re > 107) and solid walls, or in flows
involving a wide range of physical scales (such as a jet issuing into a large chamber), the
lower bound values at which clipping occurs must be lowered. This is done using the
CLIP keyword on the VISCOSITY command by assigning a value to this keyword that is
greater than 107.
Note that streamline upwinding and clipping should be regarded as being artificial stabil-
ity-enhancing measures, because they interfere with the course of the numerical solution.
This is especially true of streamline upwinding. If used correctly, the amount of upwind-
ing used must be just enough to suppress anomalous stream-wise oscillations. Excessive
upwinding (accomplished using larger than required upwinding factors) can lead to false
numerical diffusion in the stream-wise direction which can in turn modify the true
numerical solution. The default upwinding factor employed in FIDAP should be optimal
for most simulations.

© Fluent Inc., Dec-98 7-27


Checklists MODELING TURBULENT FLOWS

7.4 Checklists
7.4.1 Turbulence Modeling Checklist

The following is a checklist of the steps specific to a simulation using a turbulence model:
1. Specify the TURBULENT keyword on the FIPREP PROBLEM command.
2. Specify any turbulent constants or specialized options using the TURBOPTIONS
command.
3. Select the turbulence model desired with the MIXLENGTH, MIXING, or TWO-
EQUATION keywords on the VISCOSITY command.
a) If you specify the MIXLENGTH model, you must supply the USRMXL sub-
routine and create an executable module. (See Appendix C of the FIDAP
User’s Manual for a general discussion of user subroutines.)
b) If you specify the TWO-EQUATION model:
i) Specify a two-equation model by means of the TURBOPTIONS
command. (The default STANDARD option indicates the use of the stan-
dard k - e model.)
ii) Specify an eddy viscosity constitutive relation by means of the EDDYVIS-
COSITY command. (The default BOUSSINESQ option indicates the use
of the Boussinesq eddy viscosity constitutive relation.)
iii) Specify any desired constrained values for k and e (or w ) at the inlet and
entrainment boundaries using the KINETIC and DISSIPATION keywords,
respectively, on the BCNODE command. (Specification of k and e (or w )
is not required on wall boundaries.)
iv) Specify any desired initial values for both k and e (or w ) using the
KINETIC and DISSIPATION keywords, respectively, on the ICNODE
command. This step is strongly recommended.
4. For any solid wall boundaries, specify WALL boundary elements (see Section
7.4.2, below).

7-28 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Checklists

7.4.2 Wall Boundary Elements Checklist and Recommendations

The following sections present a checklist for the specification of wall boundary elements
and recommendations that apply to the following flow types and turbulence models:
· High-Reynolds number (high-Re) k - e models
· Low-Reynolds number (low-Re) k - w models

Wall Boundary Elements Checklist

The following is a checklist of the steps particular to the specification of wall boundary
elements for turbulent simulations:
1. Define a WALL boundary entity on any wall boundaries of the model.
2. Create the mesh:
a) For high-Re k - e models, create the mesh such that the first layer of elements
is thick enough to completely contain the viscous sublayer and transition
region in the near wall region.
b) For low-Re k - w models, use enough grid points in the near-wall regions to
accurately resolve the sharp profiles of the mean flow variables as well as the
turbulence variables.
3. Set velocity components on wall boundaries just as they would be set for non-
turbulent analyses—that is, fixed walls are defined with all velocity components
set to zero and moving walls have the appropriate velocity component set to a
specified velocity.
4. Input constants affecting the near-wall modeling on the TURBOPTIONS
command. (In almost all cases the defaults should be satisfactory.)
5. Nodes on a WALL boundary do not require the specification of any kinetic energy
or dissipation boundary conditions.

© Fluent Inc., Dec-98 7-29


Checklists MODELING TURBULENT FLOWS

Recommendations

High-Re Models

After the solution, to assure that the first layer of elements is thick enough to completely
contain the viscous sublayer and transition region, use the YPLUS command in FIPOST
to plot the y + values at the WALL boundaries. (The y + value is a dimensionless distance
related to the various layers present in a turbulent flow. You can have y + values for
momentum, temperatures and species, depending on the set of equations you are solving.)
If the value of y + for the momentum layer is greater than 30 for all elements then these
elements are thick enough. Values of y + lower than 30 for the momentum layer may be
safely tolerated provided these are not occurring in wall regions of crucial importance to
the overall flow process.
In isothermal flows, the predicted velocity field is generally insensitive to y + values in the
range 10 < y + < 1000 for most wall boundaries where the flow remains attached to the
wall. This range of y + values is even appropriate when flow separation occurs resulting
from sharp corners or edges, because such separation behavior is very pronounced and is
due primarily to inviscid mechanisms removed from the wall and not to viscous mecha-
nisms present in the viscous sublayers.
In flow problems involving subtle separation phenomena, such as separation occurring on
gently sloping surfaces (for example, flow in diffusers) or on curved surfaces (for
example, flow in turning ducts or U-bends), the predicted flow field will be sensitive to
the y + values upstream of the separation point. In these situations the most accurate pre-
dictions will be obtained if the y + values upstream of regions of potential flow separation
are kept in the range 30 < y + < 100 . If after obtaining a solution y + values in these
regions are found to be outside the above range, then a further run with a modified mesh
in the near-wall region may have to be performed. If the y + values were too large, then a
finer mesh in the near-wall region should be employed. Conversely, if y + values are less
than 30, a mesh with thicker near-wall elements should be employed.
In non-isothermal simulations involving crucial heat and mass transfer phenomena
through the walls, the allowable range of momentum y + values is narrower compared to
that of isothermal flows. In these situations it is recommended that the y + values be kept
in the range 30 < y + < 300 . If separation is also present in the flow, then for obtaining
the best possible prediction it is recommended to adhere to the even narrower range—that
is, 30 < y + < 100 .

Note that the thermal and species y + values should not be used as a guide for adjusting
the thickness of the near-wall elements. These values simply reflect the relative thickness
of the thermal and/or species sublayers compared to the viscous sublayer and have been

7-30 © Fluent Inc., Dec-98


MODELING TURBULENT FLOWS Checklists

made available for plotting in FIPOST to shed more light on the underlying physical
processes. For example, in gaseous flows where the laminar Prandtl number is of the
order of unity, the thermal y + values will be close to the momentum y + values indicating
that the viscous and thermal sublayers have similar thicknesses. In high Prandtl number
fluids such as oils where the thermal sublayer is much thinner than the momentum
sublayer, the thermal y + value will be much larger than the momentum y + values.
Conversely, in low Prandtl number fluids such as liquid metals, where due to the
enhanced thermal diffusion the thermal sublayers are very thick, the thermal y + values
will be much smaller than those of the momentum y + values. Similar considerations apply
to the relative thicknesses of the species and momentum sublayers and their correspond-
ing y + values where the relevant dimensionless number representing the relative impor-
tance of momentum to species diffusion is the laminar Schmidt number.

Low-Re Models

For low-Re k - w simulations, it is not possible to check the y + values in FIPOST (as can
be done for the special elements while using the k - e models). The best way to insure that
the near-wall mesh is sufficiently fine is to compare the value of the molecular viscosity
to the value of the turbulent viscosity, m t , on nodes adjacent to the wall. The value of m t
must not exceed m in those regions. If m t exceeds m , a finer mesh should be used in the
near-wall region.

© Fluent Inc., Dec-98 7-31


MODELING COMPRESSIBLE FLOWS Preliminary Remarks

8. MODELING COMPRESSIBLE FLOWS


FIDAP provides extensive capabilities for the simulation of compressible flows. This
chapter summarizes these capabilities and the steps required to set up and run a com-
pressible flow simulation using FIDAP. More complete details on the theoretical under-
pinnings of both the equations solved and the numerical techniques employed can be
found in the FIDAP Theory Manual.

8.1 Preliminary Remarks


As a prerequisite to reading this chapter you are strongly recommended to read Chapter 2
of the FIDAP Theory Manual which describes in detail the various flow governing equa-
tions. One such equation is the equation of state (EOS) which in its most general form
relates density, temperature, pressure and species concentrations. This relationship may be
written as;
f (r, p, T , c1 , c2 ,...) = 0 . (8.1.1)

In FIDAP the equation of state is used to define density in terms of other flow variables
and is specified via the fluid density property model. The density model is input using the
DENSITY command which is then associated with a particular entity (for example, a
fluid) via the ENTITY command.
Various types of density models may be employed in FIDAP ranging from constant
density models to models where the density is a function of pressure, temperature and
species concentration. The density model also determines whether the flow process is to
be considered as being incompressible or compressible. This is a very important distinc-
tion as it determines the exact form of flow equations that will be employed in FIDAP and
the choice of algorithms that are available for solving these equations. As noted above,
this chapter is mainly intended for describing the compressible flow capabilities of
FIDAP. However, before this is done a brief description of incompressible flow capabili-
ties is given in the following section.

© Fluent Inc., Dec-98 8-1


Incompressible Flows MODELING COMPRESSIBLE FLOWS

8.2 Incompressible Flows


The density variation in many flow processes of practical interest is negligibly small.
Examples of these are flow processes involving liquids such as water, oils, molten metals,
etc. Other examples are flow processes involving gases subject to moderate pressure
and/or temperature gradients. Such flows can be accurately modeled by assuming the
density to be a constant. The appropriate EOS is therefore an equation of the type,
r = r0 (8.2.1)

which is the simplest form of equation (8.1.1). Such an equation of state is specified in
FIDAP via the DENSITY command using the CONSTANT model, where the value of
the CONSTANT keyword is the value of r 0 . An incompressible flow is specified using
the INCOMPRESSIBLE keyword on the PROBLEM command; in this case, all density
models employed in a flow simulation must be of the CONSTANT type. Note that in an
incompressible flow simulation the density need not be the same everywhere in the com-
putational domain. An example of this is a flow problem involving two fluids (say, air and
water) separated by a solid region. Another example is a flow problem involving two
immiscible fluids (say, oil and water) separated by a sharp internal interface. In both these
examples the density of each fluid is input using a CONSTANT density model.
The most important ramification of an INCOMPRESSIBLE flow simulation is that the
continuity equation assumes the following simplified form:
Ñ×u = 0 (8.2.2)

which is often referred to as the incompressibility constraint. The density appearing in the
remainder of the flow governing equations assumes its prescribed CONSTANT value.
The only exception to this is in flows involving buoyancy forces (the BUOYANCY key-
word specified on the PROBLEM command) where the Boussinesq assumption is
employed to model the body force term (r - r 0 )g which appears in the momentum
equation. For this term only, the CONSTANT density model allows two representations
of the density variation which may result from temperature and/or species concentration
variations—but not pressure variations. These two alternative representations may be
selected by specifying the TYP1 or TYP2 keywords of the DENSITY command. For
more details on the exact forms of the buoyancy term refer to Chapter 9 of the FIPREP
User’s Manual where the DENSITY command is discussed in detail.

8-2 © Fluent Inc., Dec-98


MODELING COMPRESSIBLE FLOWS Compressible Flows

8.3 Compressible Flows


There are also many flow processes encountered in practice where the fluid density varies
significantly. This typically occurs in flow processes involving gases where the pressure
and/or temperature variations are large and of the order of the absolute pressure and/or
temperature levels. Significant density variations may also result in gaseous mixtures
involving reacting chemical species provided that the chemical reactions produce signifi-
cant changes in the gas constant of the mixture. Noticeable density variations may also
occur in liquids. However, these are not as large as typical density variations encountered
in gases and are due primarily to temperature and/or species concentration variations.
Finally, small density changes may also occur in liquids as a result of extreme pressure
variations, but these situations are not as common.
A large number of equations of state exist which approximate the dependence of density
on the other primary flow variables (see equation (8.1.1)). Some of these are more general
and can be applied to a large class of fluids under a relatively large range of working con-
ditions. Others are more specialized, intended for specific fluids operating under a
narrow, and often extreme, range of working conditions. FIDAP offers three main non-
constant categories of density models from which the user may construct a relatively wide
range of equations of state. These are:
· Variable density model
· Ideal gas law density model
· User-specified density model entered by means of a user-supplied subroutine
These density models are respectively activated via the VARIABLE, IDEAL and SUB-
ROUTINE keywords of the DENSITY command. Before these non-constant density
models are discussed in detail it is necessary to introduce a number of definitions, con-
ventions and notions that are of fundamental importance to FIDAP and its proper opera-
tion. While not all of these are unique to the modeling of compressible flows, they are
particularly relevant to such simulations.

8.3.1 Definitions and Conventions

This section defines the types of flows that FIDAP considers as being compressible. It
provides definitions for single- and multi-component flows, and introduces the notion of
the carrier fluid as it relates to single- and multi-component fluids.
When a compressible flow is specified by the COMPRESSIBLE keyword on the PROB-
LEM command, then at least one of the density models input must be of the non-constant
type (that is, VARIABLE, IDEAL or SUBROUTINE). It is also very important to note
that there is a fundamental difference in the types of equations of state that can be con-
structed with the above three types of density models. The VARIABLE density model

© Fluent Inc., Dec-98 8-3


Compressible Flows MODELING COMPRESSIBLE FLOWS

allows the density to be a function of temperature and/or species concentration, but does
not allow any dependence on pressure. In contrast, the IDEAL and SUBROUTINE
density models require as a minimum that density be a function of both pressure and tem-
perature; if chemical species are also present in the flow process, you may optionally
choose to make density a further function of species concentrations. When density is
allowed to be a function of pressure, the physical flow process becomes more intricate
and intercoupled. This gives rise to phenomena that would not otherwise be present.
Probably the most significant of these is the phenomenon of sound wave propagation. An
important consequence of this is that shocks may be produced in flow regions where the
local speed of flow exceeds the local speed of sound. Since the flow processes resulting
from the VARIABLE and IDEAL/SUBROUTINE density models are fundamentally dif-
ferent, FIDAP does not allow the mixing of these two classes of density models in the
same simulation. Thus, a COMPRESSIBLE simulation may be performed with various
density models of the VARIABLE type. Alternatively a COMPRESSIBLE simulation
may be performed with various density models of the IDEAL and/or SUBROUTINE
type. However, such a simulation may not be performed involving a VARIABLE density
model and an IDEAL (or SUBROUTINE) density model concurrently. Because of these
restrictions, the form of the energy equation adopted in FIDAP in a COMPRESSIBLE
flow simulation will depend on the particular density model employed. Additionally, if the
simulation is time-dependent (TRANSIENT keyword on the PROBLEM command), the
form of the continuity equation will depend on the solution algorithm employed. Specific
details are provided later in the following sections where the above three density models
are more fully described.
The notion of the carrier fluid is of fundamental importance to FIDAP and has a some-
what different connotation depending on whether a single- or multi-component simulation
is being performed. A single-component simulation comprises a single homogeneous
fluid. Examples of such fluids are pure substances such as water or nitrogen gas. How-
ever, in the context of FIDAP, non-pure substances may also qualify as being single-com-
ponent fluids provided they are well mixed. The best known example of this is air which
is composed of a homogeneous mixture of pure gases. In a single-component simulation
the working fluid is the carrier fluid. This means that the various flow governing equa-
tions of momentum, continuity, temperature, etc. (no species equations may of course be
present in this context), and all property models defined in FIDAP (that is, for specific
heats, conductivities, etc., as well as densities), pertain to the carrier fluid.
A multi-component simulation comprises fluids involving a non-homogeneous mixture of
more that one substance. The convention used in FIDAP to determine the number of
components N present in the mixture is as follows. The number of components is equal to
the number of species equations present in the flow, plus one—the carrier fluid. For
example, in a flow simulation involving a three-component mixture (N=3), only two
transport equations need to be solved (for c1 and c2 )—the species concentrations of any
two of the three mixture components. The third component is regarded as being the carrier

8-4 © Fluent Inc., Dec-98


MODELING COMPRESSIBLE FLOWS Compressible Flows

fluid which is arbitrarily chosen to be any one of the three components of the mixture. In
a multi-component mixture the various flow governing equations (momentum, continuity,
temperature, species concentration, etc.) and all property models defined in FIDAP
pertain to the mixture. Moreover, the principle of conservation of matter dictates the
following relationship between the species concentration cN of the carrier fluid and the
species concentrations of the remaining N-1 species components;
N -1
cN = 1 - å cn . (8.3.1)
n=1

This relationship is inherent in all the non-constant density models available in FIDAP
and provides the crucial link which relates the density of the mixture to the mass fractions
of the various chemical components present in the flow. It is important to note that in its
strictest sense the summation in equation (8.3.1) involves all the N-1 chemical species
present in the flow simulation as it is assumed that all these will contribute to the density
of the mixture. However, in practice this is not always the case as there are situations
where some of the species components may be neutral (or near neutral) from the stand-
point of the overall mass of the mixture. Examples of these are tracer fluids which may or
may not be chemically inert. For this reason when a non-constant DENSITY model in
entered in FIDAP, you are required to explicitly specify which chemical species contrib-
ute to the summation in equation (8.3.1). The manner in which this is done is explained in
the following section.

8.3.2 The VARIABLE Density Model

The VARIABLE density model allows the density to be represented in the general form:

b
r 0 1 - b T T - T0 g
r=
F
1+ å G
N -1
M I
- 1J c
.
(8.3.2)
HM K
N
n
n=1 n

The following approximate version of equation (8.3.2) is also available which is referred
to as the product form:

b g Õ d1 - b c i
N -1
r = r 0 1 - b T T - T0 cn n . (8.3.3)
n

The above forms of the VARIABLE density model are invoked via the TYP1 or TYP2
keywords of the DENSITY command. TYP1, which is the default form, selects equation
(8.3.2) and TYP2 selects equation (8.3.3). In the above equation, r 0 is the reference
density which is the value of the VARIABLE keyword. b T and b c are respectively, the
coefficients of volume expansion due to temperature and species concentration, and T0 is
a reference temperature. The value of T0 and the property models for b T and b c are

© Fluent Inc., Dec-98 8-5


Compressible Flows MODELING COMPRESSIBLE FLOWS

entered via the VOLUMEXPANSION command (see Chapter 9 of the FIPREP User’s
Manual for details). Mn is the molecular weight of the n’th chemical species, and M N is
the molecular weight of the carrier fluid. The molecular weights are entered via the
MOLECULARWEIGHTS command (see Chapter 9 of the FIPREP User’s Manual for
details).
Recall that in setting up a density model, you are required to specify explicitly which of
the species components present in the simulation will be contributing to the summations
present in the above equations. How this is done is demonstrated by way of the following
example. Consider a six-component COMPRESSIBLE flow simulation involving the five
species, c1, c2 , c3 , c4 , and c5 (that is, keywords SPECIES=1, SPECIES=2, ..., SPECIES
=5 specified on the PROBLEM command), where you wish to define a density model
making density a function of the three species components c2 , c3 , and c5 only. The
DENSITY command for that particular model must include the three keywords,
SPECIES=2, SPECIES=3 and SPECIES=5. If these keywords are not specified the
density will not be a function of species concentration.
Equations (8.3.2) and (8.3.3) above may be manipulated to create a wide range of equa-
tions of state for density.
In the case of a single-component fluid (carrier fluid only), the VARIABLE model can
only be used if the temperature equation is present (ENERGY or SPECIES=0 keyword on
PROBLEM command). In this context the form of equations (8.3.2) and (8.3.3) are the
same and is as follows:

b
r = r 0 1 - b T T - T0 g . (8.3.4)

In the case of a multi-component fluid, if the temperature equation is not present (or if it
is present but the TEMPERATURE keyword is not specified on the DENSITY com-
mand) the VARIABLE density model assumes different forms depending on whether the
TYP1 or TYP2 keywords are specified. If TYP1 is specified, the following form is
employed:
r0
r=
FM
1+ åG
N -1
I
- 1J c
. (8.3.5)
HM K
N
n
n= 1 n

If TYP2 is specified the density equation becomes:

8-6 © Fluent Inc., Dec-98


MODELING COMPRESSIBLE FLOWS Compressible Flows

d i
N -1
r = r 0 Õ 1 - b c cn
n
. (8.3.6)
n

Again remember that it is your responsibility to ensure which species components are
activated in the above expressions by selecting the appropriate SPECIES=n keywords on
the DENSITY command.

© Fluent Inc., Dec-98 8-7


Compressible Flows MODELING COMPRESSIBLE FLOWS

NOTES
1. The VARIABLE and IDEAL (or SUBROUTINE) density models cannot be used
together in a COMPRESSIBLE simulation.
2. In a COMPRESSIBLE flow simulation involving VARIABLE type density mod-
els, the form of the temperature equation employed in FIDAP is as follows:

F ¶T + u ¶T I = - ¶q
rc p GH ¶t ¶x JK ¶x
j
j
j

j
+H (8.3.7)

where c p is the specific heat at constant pressure, q j is the flux of thermal energy
and H is the general heat generation/destruction term.
3. The form of the continuity equation in TRANSIENT flows will depend on the
type of solution algorithm employed. If a SEGREGATED solution algorithm is
employed, the full form of the continuity equation is employed—that is,

c h
¶r ¶ ru j
+ =0 . (8.3.8)
¶t ¶x j
If one of the fully coupled solution algorithms is employed (that is, either one of
the S.S., N.R., Q.N. or M.N. keywords on the SOLUTION command), the fol-
lowing truncated form of the continuity equation is employed:

c h
¶ ru j
=0 . (8.3.9)
¶x j
Moreover, if a fully coupled solution algorithm is employed, only a MIXED pres-
sure-velocity formulation is allowed—the PENALTY formulation may not be
employed. Note finally that a TRANSIENT COMPRESSIBLE simulation may not
be performed with the explicit FORWARD Euler solution algorithm.

8-8 © Fluent Inc., Dec-98


MODELING COMPRESSIBLE FLOWS Compressible Flows

8.3.3 The IDEAL and SUBROUTINE Density Models

Recall from above that the IDEAL and SUBROUTINE density models require that den-
sity is at least a function of temperature and pressure. This means that in COMPRESSI-
BLE flow simulations involving IDEAL and/or SUBROUTINE density models at least
the momentum, continuity and temperature equations must be present (the ENERGY (or
SPECIES=0) and MOMENTUM keywords specified on the PROBLEM command).
The IDEAL density model activates the well-known ideal gas law equation of:
MN p
r=
L FM
RT M1 + å G
N -1
I O
- 1J c P
.
(8.3.10)
N HM K Q
N
n
n= 1 n

In the above equation, p and T are the absolute thermodynamic pressure and temperature
respectively, and R is the universal gas constant. The value of the universal gas constant
R may be entered in one of two ways. A unique constant value may be entered for each
fluid ENTITY present in the computational domain via the GASCONSTANT keyword on
the DENSITY command. If, on the other hand, R is constant throughout the computa-
tional domain (which is usually the case), its value may be entered as a global parameter
via the GASCONSTANT keyword on the COMPRESSIBLE command. It is important to
note that the global value of R entered using the GASCONSTANT keyword of the
COMPRESSIBLE command will be overridden by any R value entered via the GAS-
CONSTANT keyword of the DENSITY command.
Note the similarity between the above equation and equation (8.3.2) of the VARIABLE
density model which has been derived from the ideal gas law model. The significant dif-
ference between them is the appearance of pressure in equation (8.3.10). A further differ-
ence is that the temperature appearing in the VARIABLE density model need not
necessarily be the absolute thermodynamic temperature—it may be the temperature per-
turbation about a reference temperature T0 .

Finally, as is the case in the VARIABLE density models, it is your responsibility to en-
sure which species components are active in a given IDEAL density model by selecting
the appropriate SPECIES=n keyword(s) on the DENSITY command.
The SUBROUTINE density model allows you to construct “ideal-gas-law-type” equa-
tions of state of the form:
1 FG p IJ
r
=f
HTK . (8.3.11)

© Fluent Inc., Dec-98 8-9


Compressible Flows MODELING COMPRESSIBLE FLOWS

For this option, you must supply a subroutine USREOS that returns the value of the
density at any point in the fluid based on such an equation of state. The format of the
USREOS subroutine is presented in Chapter 9 of the FIPREP User’s Manual where the
DENSITY command in described. Although a great deal of flexibility for constructing
customized equations of state is provided by the USREOS subroutine, the equation of
state constructed in this manner must satisfy the functional form of equation (8.3.11).

NOTES
1. As noted above, the phenomenon of sound wave propagation is present in COM-
PRESSIBLE flow simulations employing the IDEAL and/or SUBROUTINE den-
sity models. This means that shock waves may potentially be formed in the flow if
the fluid speed is large. FIDAP is only intended for flows where the fluid speed is
lower than the speed of sound—that is subsonic flows where the Mach number in
the flow is everywhere less that one (Ma < 1).
2. In a COMPRESSIBLE flow simulation employing IDEAL and/or SUBROUTINE
density models, the form of the energy equation solved in FIDAP is:
F ¶T + u ¶T I = - ¶q ¶u j
rcv GH ¶t ¶x JK ¶x
j
j
j

j
+ H - Ec p
¶x j
(8.3.12)

where cv is the specific heat at constant volume whose value is computed in each
fluid ENTITY from the following equation:

R LM
N -1
1+ å
FG
MN IJ OP
cv = c p -
MN N H
n = 1 Mn
- 1 cn
K Q . (8.3.13)

Note that no additional property model need be entered for computing cu as all
information necessary for its evaluation is already available through the property
models entered for the fluid entity for c p , R and the molecular weights, Mn and
M N . Also, the species components active in the summation in equation (8.3.13)
are the same as those that have been activated in the DENSITY model entered for
the fluid entity.
The last term in equation (8.3.12) is the reversible pressure work term which
causes significant temperature changes in flows where Ma > 0.3. The dimension-
less coefficient Ec multiplying the pressure work term is provided to facilitate the
nondimensionalization of the energy equation. This coefficient is assigned to a
particular fluid entity via the REVERSIBLE keyword of the ENTITY command.
If the flow problem is solved in dimensional form, the value of the coefficient Ec
is unity which is the default value of the REVERSIBLE keyword. However, if the
flow problem is formulated and solved in dimensionless form, the value of Ec

8-10 © Fluent Inc., Dec-98


MODELING COMPRESSIBLE FLOWS Compressible Flows

may not be unity, and therefore needs to be explicitly entered using the
REVERSIBLE keyword.
In a COMPRESSIBLE simulation using the IDEAL and/or SUBROUTINE den-
sity models, the flow problem must always be set up in terms of absolute tem-
peratures. It is strongly recommended that a non-zero initial (or guess) temperature
field be prescribed everywhere in the computational domain. This can be done
through the use of the ICNODE(TEMP, CONSTANT = v, ALL) command.
3. The correct modeling of pressure is also of crucial importance in flows employing
the IDEAL and/or SUBROUTINE density models. In these simulations the level
of absolute pressure must be correctly imposed and strictly controlled. (This con-
trasts with flow simulations employing CONSTANT or VARIABLE density mod-
els where the absolute level of pressure is unimportant and it is only necessary to
model pressure variations with respect to an arbitrarily set pressure datum.) The
absolute pressure level may be set by a combination of the normal stress boundary
condition applied at an outflow boundary and a reference pressure level value
imposed through the PRESSURE keyword of the COMPRESSIBLE command. It
is strongly recommended that a non-zero reference pressure level be set. All com-
puted pressures in FIDAP will then be perturbations about this reference value,
where also the absolute pressure will be taken as being the computed pressure
perturbation plus the reference value.
Consider for example a flow problem where the absolute pressure pexit is known
at the exit plane of the computational domain. This value of pressure may be
imposed in the numerical solution by a combination of setting the value of the
PRESSURE keyword to pexit and prescribing a zero normal traction boundary
condition at the outflow boundary. This outflow boundary condition sets the value
of the computed pressure to zero at the exit plane. However, when this is added to
the reference value of pexit the required absolute level of pressure is recovered. If
a further flow exit plane exists where the absolute pressure is different from pexit
by an amount Dp, then the correct level of pressure is set there via a non-zero
normal traction boundary condition which has a value of Dp. This non-zero value
of traction is entered via the BCFLUX command.
Note finally that fully confined COMPRESSIBLE flows employing the IDEAL
and/or SUBROUTINE density models cannot be modeled with the current version
of FIDAP. This is because there is no mechanism currently available through
which to relate the absolute pressure level to the total amount of mass present in
the domain. This mechanism will be made available in future versions of FIDAP
and it will involve a mass constraint condition similar to the volume constraint
condition which is available for free surface simulations in FIDAP.

© Fluent Inc., Dec-98 8-11


Compressible Flows MODELING COMPRESSIBLE FLOWS

4. Only the SEGREGATED solution algorithm is available for COMPRESSIBLE


flow simulations involving the IDEAL and/or SUBROUTINE density models.
Moreover only continuous pressure interpolations are allowed (CONTINUOUS
keyword on the PRESSURE command). It is also strongly recommended for
elements with quadratic interpolation functions.
5. It is recommended that larger relaxation factors for pressure be used compared to
flow simulations involving CONSTANT or VARIABLE density models. The
degree of pressure relaxation depends on the maximum Mach number encountered
in the flow. For Ma < 0.3, the default relaxation factor employed by FIDAP is
often adequate. For 0.3 < Ma < 0.85, pressure relaxation factors of up to 0.85 may
be needed. Flows involving regions where the flow conditions are just below (or
possibly just above) sonic conditions may require relaxation factors as large as
0.95.
6. In TRANSIENT COMPRESSIBLE flow simulations employing the IDEAL
and/or SUBROUTINE density models, the full form of the continuity equation is
employed—that is,

+
c h
¶r ¶ ru j
=0 .
¶t ¶x j
7. The value of the IDEAL keyword is the reference density r 0 . This value only
needs to be entered in a COMPRESSIBLE flow simulation if a buoyancy force
b g
r - r 0 g is present in the momentum equation (BUOYANCY keyword on the
PROBLEM command).
8. The DELAY keyword on the COMPRESSIBLE command provides control on the
numerical stability of the solution in difficult flow problems where Ma » 1. This
keyword is used to postpone the activation of various terms in the flow equations
which cause flow compression for a given number of user-specified iterations n.
The temperature equation is also heavily relaxed for the first n iterations. The
value of the DELAY keyword is the number of initial iterations n. This procedure
allows the flow field to stabilize before introducing flow compression effects and
tends to stabilize the convergence process. The default value of this keyword is 5.
For difficult problems, larger values (for example, 15) may be used. If you are
specifying an iteration strategy (SEGREGATED keyword on the STRATEGY
command), n will be the number of solutions of the pressure equation.

8-12 © Fluent Inc., Dec-98


MOVING-BODY CONSTRAINT CAPABILITY Introduction

9. MOVING-BODY CONSTRAINT CAPABILITY


9.1 Introduction
FIDAP includes a capability which allows the user to define constraints that are applied
and released as a function of time and position for the various nodal degrees of freedom.
One application of such constraints is the modeling of a solid body moving through a
continuum of fluid; therefore, the new capability is referred to as the “moving-body con-
straint” capability.
A typical application for the moving-body constraint capability is the modeling of a
stirred, baffled tank such as that shown schematically in Figure 9-1, in which the velocity
along the walls and baffles is zero, and the velocity along the blades is given by the rota-
tional speed. At any given time step, the nodes associated with the blade may be given
their appropriate boundary conditions or constraints. At the next time step, however, the
blade is in a different position, therefore the previous constraints must be removed and
new constraints applied. The moving-body constraint capability may be used in such
instances to apply and remove constraints in a systematic fashion.

wall

baffles
blades

Figure 9-1: Schematic of stirred tank with baffles

In addition to constraints based on position, time and/or solution, any given nodal degree
of freedom (for example, temperature) may be constrained based on the value of other
degrees of freedom (for example, velocity). As a result, the moving-body capability
allows the construction of complex, solution-dependent nodal constraints which may be
used in either steady-state or transient simulations.

© Fluent Inc., Dec-98 9-1


Usage MOVING-BODY CONSTRAINT CAPABILITY

9.2 Usage
In many ways, the moving-body constraint capability in FIDAP is similar to the standard
FIDAP BCNODE command. In both cases, a particular degree of freedom at a given node
is constrained to a prescribed value, which may be a function of time, position and other
solution variables. They differ, however, in that the BCNODE-type constraint applies to
the nodal degree of freedom for the entire simulation. A moving-body constraint, on the
other hand, is typically applied and released as a function of time and position, the sched-
ule of which is defined in a user subroutine (described below).
The moving-body constraint capability in FIDAP is intended to supplement, rather than
replace, the BCNODE command. The BCNODE command defines the constraint condi-
tion prior to the main computation, and the BCNODE constraints are defined only once
for the particular run (even though they may be time, position and/or solution dependent).
By contrast, the moving-body constraint capability defines and applies constraints that
may be released (that is, not constrained) during the main computational part of the run.
The moving-body constraint, therefore, should only be used for constraints which need to
be alternatively applied and released during the same run.
As is the case for the BCNODE command, the moving-body constraint can be used to
apply conditions to the following degrees of freedom:
· x component of velocity, ux
· y component of velocity, uy
· z component of velocity, uz
· Temperature, T
· Free surface position, s
· Turbulent kinetic energy, k
· Turbulent dissipation, e
· Species 1 through 13

Constraints cannot be applied to PRESSURE, SPECIES = 14, or SPECIES = 15. (NOTE:


SPECIES 14 and 15 are reserved for the moving-body constraint and the filling capabili-
ties, respectively.) The free-surface degree of freedom can be used with the moving-body
constraint capability only when the segregated solver is employed.

9-2 © Fluent Inc., Dec-98


MOVING-BODY CONSTRAINT CAPABILITY Usage

It is important to note that flux-type boundary conditions are overridden by the moving-
body constraint. Such flux conditions include those specified through the BCFLUX
command, as well as those determined for a CONVECTION or ESPECIES entity. Con-
straints defined using the BCNODE command, however, override the moving-body
constraint.
Special care must be taken when applying moving-body constraints to nodes on periodic
boundaries. If a moving-body constraint is applied to a node on the reference face, then an
equivalent constraint must also be applied on the periodic node (and vice versa). A con-
straint applied only to a reference node or to a periodic node will produce unexpected
results.

9.2.1 Activating the Moving-Boundary Constraint

The moving-body constraint is activated in FIPREP on an ENTITY basis. The ENTITY


command,
ENTITY( NAME = "s", ..., MOVING )
activates the moving-body constraint for the nodes in that particular entity. For the sake of
efficiency, only those entities including nodes to which the moving-body constraints are
to be applied should be flagged with the MOVING keyword. If no ENTITY is defined as
MOVING, then no moving-body constraints are applied.
Any type of continuum or boundary entity may be flagged with the MOVING keyword.
During the solution phase, a user subroutine, USRMBC, is then called for each node
which belongs to at least one ENTITY flagged with the MOVING keyword. The subrou-
tine returns information to the solver regarding whether the solution for a particular
degree of freedom is to be solved for or constrained. If the solution is to be constrained,
the user subroutine also returns the constrained value.
The moving-body constraint requires the preparation of a user-supplied subroutine,
USRMBC. The form of the subroutine (shown below) is similar to subroutine USRBCN,
which defines a BCNODE-type boundary condition.
SUBROUTINE USRMBC (VAL, NODE, TIME, SOL, ID, NDOF, NUMNP, LDOFU,
1 CONSTR, NODEPR, XYZ, IFLAG, IERR)
C
C INPUT:
C NODE - INTERNAL NODE NUMBER
C TIME - VALUE OF TIME COUNTER
C SOL - SOLUTION VECTOR
C ID - GLOBAL EQUATION NUMBER ARRAY;
C NDOF - NUMBER OF ACTIVE DEGREES OF FREEDOM
C NUMNP - NUMBER OF NODES IN MESH
C LDOFU - ACTIVE DEGREE OF FREEDOM ARRAY
C CONSTR - ARRAY OF SPECIFIED DEGREE OF FREEDOM VALUES

© Fluent Inc., Dec-98 9-3


Usage MOVING-BODY CONSTRAINT CAPABILITY

C NODEPR - INTERNAL TO EXTERNAL NODE TRANSLATION ARRAY


C XYZ - ARRAY OF NODAL COORDINATES; DIMENSION NUMNP X 2/3
C
C OUTPUT:
C VAL - ARRAY OF IMPOSED VALUES FOR DOFS
C IFLAG - ARRAY OF FLAGS FOR DOFS
C
#include "IMPLCT.COM"
#include "PARUSR.COM"
DIMENSION SOL(*),ID(NUMNP,NDOF),LDOFU(*),XYZ(NUMNP,*)
DIMENSION IFLAG(*),VAL(*),NODEPR(*),CONSTR(*)
C
ZRO = 0.D0
C
C .... BODY OF SUBROUTINE ....
C
IFLAG(KDU) = ....
IFLAG(KDV) = ....
.
.
.
VAL(KDU) = ....
VAL(KDV) = ....
.
.
.
C
RETURN
END

Inputs to the subroutine USRMBC include: the current time, an internal node number, the
nodal coordinates and the nodal solution. The subroutine outputs two arrays: IFLAG and
VAL. The IFLAG array specifies which degrees of freedom are to be constrained for a
given node. Its possible indices are:

NDF Index Degree of Freedom

KDU x component of velocity, ux

KDV y component of velocity, uy

KDW z component of velocity, uz

KDK turbulent kinetic energy, k

KDE turbulent dissipation, e

9-4 © Fluent Inc., Dec-98


MOVING-BODY CONSTRAINT CAPABILITY Usage

NDF Index Degree of Freedom

KDF free-surface position, s

KDT temperature, T

KDS + n species = n, n = 1, 2, ..., 13

A particular degree of freedom, NDF, for the input node is constrained if the associated
value of IFLAG(NDF) is not equal to zero. If IFLAG(KDU) = 1, for instance, then the
degree of freedom UX will be constrained for that node. The VAL array provides the
value of the constrained degree of freedom for a given node.
The constrained value for degree of freedom NDF for the node is defined by VAL(NDF).
The indices of the VAL array follow those for IFLAG in the USRMBC subroutine. It is
only necessary to define the nonzero values of VAL. For example, the statement
IFLAG(KDT) = 1, by itself, is a correct and acceptable way to set the temperature at a
particular node to zero.
If a node belongs to a SLIP or SREACTION entity, the moving-body constraint must be
applied with respect to the local normal-tangential system for that node. For a two-dimen-
sional problem, UX and UY are regarded as the tangential and normal components, UT
and UN, respectively. For three-dimensional problems, UX, UY and UZ reference UT1,
UT2, and UN3, respectively. Thus, for a node belonging to a SLIP entity in a two-dimen-
sional problem, the normal velocity component is specified by IFLAG(KDV) = 1 and
VAL(KDV) = value.
The solution vector is provided to the user subroutine USRMBC so that nodal values of
any active degree of freedom can be queried at the node NODE or at any other node in the
mesh. For example, the value of the temperature degree of freedom, KDT, at node NODE
can be evaluated using the program segment,
NADOF = LDOFU(KDT)
IF (NADOF .EQ. 0) THEN
ERROR - DEGREE OF FREEDOM IS NOT ACTIVE FOR THIS PROBLEM
ENDIF
IEQ = ID(NODE,NADOF)
IF (IEQ .LT. 0) THEN
VALUE = CONSTR(-IEQ)
ELSE IF (IEQ .EQ. 0) THEN
VALUE = 0.
ELSE
VALUE = SOL(IEQ)
ENDIF

© Fluent Inc., Dec-98 9-5


Usage MOVING-BODY CONSTRAINT CAPABILITY

The node NODE passed to the subroutine always refers to the internal node number used
in FISOLV. In FIPOST, external node numbers are displayed. The NODEPR array
allows the conversion from an internal node number to an external number as follows:
NEXT = NODEPR(NODE)

9.2.2 Postprocessing

As noted above, SPECIES = 14 is reserved for the moving-body constraint capability. It is


used in postprocessing to visualize the node or group of nodes affected by the constraint.
The nodal value of SPECIES = 14 is unity if a moving-body constraint is applied to any
active degree of freedom for that node. Otherwise, it has a value of zero. A contour value
of SPECIES = 14 can then be used to verify the locations of the applied moving-body
constraints. The FILL command in FIPOST can be used to automatically superimpose the
contour of SPECIES = 14 onto various types of plots.

9-6 © Fluent Inc., Dec-98


MOVING-BODY CONSTRAINT CAPABILITY Examples

9.3 Examples
9.3.1 2-D Stirred Tank with Baffles

Consider a mixing tank such as that shown in Figure 9-1 containing a fluid with the prop-
erties of water. The tank is 2 meters in diameter and includes four blades of radius 0.5
meters rotating at a speed of 1.047 rads/s (approximately 10 rpm). It also includes four
baffles having a thickness of 0.1 meters. The walls and baffles receive no-slip boundary
conditions through the standard application of the BCNODE command. The moving
blades are modeled using a moving-body constraint.
Figure 9-2 shows the two-dimensional mesh for the situation described above. The mesh
is composed of two continuum entities: an “inner” entity, swept by the rotation of the
blades, and an “outer” entity, completing the mesh (see Figure 9-3.). Only the inner entity
is flagged as MOVING. At each time step, it is intended that the blades coincide with the
nodal points of the mesh, therefore the mesh has a fan-type topology in the region swept
by the blades. A fixed time step is used to ensure that the moving-body constraints are
placed on element boundaries for each time step. (This approach is recommended but not
required under the moving-body constraint.)

Figure 9-2: Example 1—Stirred-tank mesh

© Fluent Inc., Dec-98 9-7


Examples MOVING-BODY CONSTRAINT CAPABILITY

inner outer

Figure 9-3: Example 1—Continuum entities

Table 9.1 shows the user subroutine USRMBC for this example. At a given value of time,
the four blades are located based upon their spatial (x,y) location. The variables XANG and
YANG represent the intersection of the particular blade with the unit circle. The variables
XV and YV represent the normalized coordinates of the node. The node is considered part
of the blade if the normalized distance between (XV,YV) and (XANG,YANG) is less than a
prescribed tolerance of 0.02. The velocity on the blades is defined as
(ux , uy ) = (-wy, wx) ,where w is the rotation speed.

Table 9.2 shows the FIPREP part of the problem setup. The key points to note are the use
of upwinding to address the convective nature of the flow equations and a fixed time step.
The Reynolds number is about 104, based upon the blade tip speed.
The NOPREDICTION keyword is used in the TIMEINTEGRATION command, thus
specifying that the solution from the n th time step be used as the prediction for time step
n + 1. The predicted solution is actually further away from the corrected solution than is
the previous solution. Absence of the NOPREDICTION keyword increases the number of
nonlinear iterations required for each time step. The use of the NOPREDICTION key-
word is generally recommended for transient moving-body constraint simulations.
Figure 9-4, Figure 9-5, and Figure 9-6 show the results for Example 1. The FIPOST file
used to generate the figures is shown in Table 9.3. The FILL command is used to outline
the shape of the moving blade within the mesh. A contour value of 0.9 is chosen in order
to keep the outline visually thin.

9-8 © Fluent Inc., Dec-98


MOVING-BODY CONSTRAINT CAPABILITY Examples

Table 9.1: Example 1—User subroutine USRMBC

SUBROUTINE USRMBC (VAL, NODE, TIME, SOL, ID, NDOF, NUMNP, LDOFU,
1 CONSTR, NODEPR, XYZ, IFLAG, IERR)
C
C INPUT:
C NODE - INTERNAL NODE NUMBER
C TIME - VALUE OF TIME COUNTER
C SOL - SOLUTION VECTOR
C ID - GLOBAL EQUATION NUMBER ARRAY
C NDOF - NUMBER OF ACTIVE DEGREES OF FREEDOM
C NUMNP - NUMBER OF NODES IN MESH
C LDOFU - ACTIVE DEGREE OF FREEDOM ARRAY
C CONTR - ARRAY OF SPECIFIED DEGREE OF FREEDOM VALUES
C NODEPR - INTERNAL TO EXTERNAL NODE TRANSLATION ARRAY
C XYZ - ARRAY OF NODAL COORDINATES
C
C OUTPUT:
C VAL - ARRAY OF IMPOSED VALUES FOR DOFS
C IFLAG - ARRAY OF FLAGS FOR DOFS
C
#include "IMPLCT.COM"
#include "PARUSR.COM"
DIMENSION SOL(*),ID(NUMNP,NDOF),LDOFU(*),XYZ(NUMNP,*)
DIMENSION IFLAG(*),VAL(*),NODEPR(*),CONSTR(*)
C
ZRO = 0.D0
C
C RETRIEVE NODAL COORDINATES:
C
X = XYZ(NODE,1)
Y = XYZ(NODE,2)
C
C INITIAL CALCULATIONS:
C
VV = SQRT(X**2+Y**2)
IF (VV .LT. 1.0D-8) RETURN
XV = X / VV
YV = Y / VV
C
THICK = 0.02D0
OMEGA = 1.047D0
ANGLE = OMEGA * TIME
PI = 3.14159265359D0
C
C LOOP OVER THE FOUR BLADES:
C
DO 10 I=1,4
C
XANG = COS(ANGLE+FLOAT(I-1)*PI/2.0D0)
YANG = SIN(ANGLE+FLOAT(I-1)*PI/2.0D0)

© Fluent Inc., Dec-98 9-9


Examples MOVING-BODY CONSTRAINT CAPABILITY

DIST = SQRT( (XV-XANG)**2 + (YV-YANG)**2 ) * VV


C
C IF THE NODE COINCIDES WITH THE BLADE:
C
IF (DIST .LT. THICK) THEN
C
C SET FLAGS AND VALUES FOR THE TWO
C VELOCITY COMPONENTS:
C
IFLAG(KDU) = 1
IFLAG(KDV) = 1
C
VAL(KDU) = - OMEGA * Y
VAL(KDV) = OMEGA * X
C
ENDIF
C
10 CONTINUE
C
RETURN
END

9-10 © Fluent Inc., Dec-98


MOVING-BODY CONSTRAINT CAPABILITY Examples

Table 9.2: Example 1—FIPREP file

FIPREP
PROBLEM(2-D,NONLINEAR,TRANSIENT)
EXECUTION(NEWJOB)
OPTIONS(UPWINDING)
/
PRINTOUT(NONE)
TIMEINTEGRATION(BACKWARD,NSTEPS=10,DT=1.87535378079E-
1,FIXED,NOPRED)
/
DENSITY(CONSTANT=1000)
VISCOSITY(CONSTANT=0.02)
/
ENTITY(TYPE="INNER",FLUID,MOVING)
ENTITY(TYPE="OUTER",FLUID)
ENTITY(TYPE="BAFFLE",PLOT)
ENTITY(TYPE="WALL",PLOT)
/
BCNODE(VELOCITY,ZERO,ENTITY="WALL")
BCNODE(VELOCITY,ZERO,ENTITY="BAFFLE")
/
END
CREATE(FISOLV)
END

Table 9.3: Examples 1 and 2—FIPOST file

FIPOST
/
FILL(ON,SPECIES=14,INPUT=1)
0.9
/
TIME(STEP=7)
CONTOUR(STREAMLINE)
/
TIME(STEP=8)
CONTOUR(STREAMLINE)
/
TIME(STEP=9)
CONTOUR(STREAMLINE)
/
END

© Fluent Inc., Dec-98 9-11


Examples MOVING-BODY CONSTRAINT CAPABILITY

Figure 9-4: Example 1—Streamline contour plot, Time = 1.31

Figure 9-5: Example 1—Streamline contour plot, Time = 1.50

9-12 © Fluent Inc., Dec-98


MOVING-BODY CONSTRAINT CAPABILITY Examples

Figure 9-6: Example 1—Streamline contour plot, Time = 1.69

© Fluent Inc., Dec-98 9-13


Examples MOVING-BODY CONSTRAINT CAPABILITY

9.3.2 2-D Baffled, Stirred Tank with Periodic Boundary Conditions

Consider again the baffled, stirred tank shown in Figure 9-1. Figure 9-7 shows one quar-
ter of the original mesh. Periodic boundary conditions for normal and tangential velocity
components are applied between the mesh boundaries defined by the lines x = 0 and
y = 0. The horizontal boundary is chosen as the reference face, while the vertical bound-
ary is the periodic face. (The choice is arbitrary in this case.)
Table 9.4 shows the user subroutine, USRMBC, for this example. The structure of the
subroutine is similar to that of the previous example but differs in two main respects.
First, in the previous case, the node at the origin is identified by its location (that is, VV
.LT. 1.0D-8); whereas, in this example, the node is identified by its node number. A
BCNODE command is used to constrain the velocity to zero at this node. Second, in this
example, nodes on the lines x = 0 and y = 0 belong to SLIP groups. The constraints
defined in the user subroutine for these nodes must be in terms for the normal/tangential
local system. The desired conditions on the reference and periodic faces become
(ut , un ) = (0,-wd) , where d is the distance from the origin. This constraint results in the
equivalent velocity vector when applied from either the reference face or the periodic
face.
The FIPREP problem setup for Example 2 is shown in Table 9.5. (Note the application of
the periodic boundary condition.) In FIDAP, the normal vector at any given face points
outward from the fluid domain, by default. Thus, for the example represented by Figure 9-
7, the normal vectors along the reference face point downward, and the normal vectors on
the periodic face point to the left. However, because the UN and UT keywords are speci-
fied on the BCPERIODIC command, the normal vector on the periodic face is effectively
changed to point inward (to the right).
Results for this example are shown in Figure 9-8, Figure 9-9, and Figure 9-10. The
FIPOST file used to generate these plots is the same as that for Example 1 (Table 9.3).

9-14 © Fluent Inc., Dec-98


MOVING-BODY CONSTRAINT CAPABILITY Examples

Figure 9-7: Example 2—Stirred-tank mesh (periodic conditions)

© Fluent Inc., Dec-98 9-15


Examples MOVING-BODY CONSTRAINT CAPABILITY

Table 9.4: Example 2—USRMBC subroutine

SUBROUTINE USRMBC (VAL, NODE, TIME, SOL, ID, NDOF, NUMNP, LDOFU,
1 CONSTR, NODEPR, XYZ, IFLAG, IERR)
C
#include "IMPLCT.COM"
#include "PARUSR.COM"
DIMENSION SOL(*),ID(NUMNP,NDOF),LDOFU(*),XYZ(NUMNP,*)
DIMENSION IFLAG(*),VAL(*),NODEPR(*),CONSTR(*)
C
ZRO = 0.D0
C
C EXCLUDE NODE ON AXIS
C
IF (NODEPR(NODE) .EQ. 183) RETURN
C
C RETRIEVE NODAL COORDINATES:
C
X = XYZ(NODE,1)
Y = XYZ(NODE,2)
C
C INITIAL CALCULATIONS:
C
VV = SQRT(X**2+Y**2)
XV = X/VV
YV = Y/VV
C
THICK = 0.002D0
OMEGA = 1.047D0
ANGLE = OMEGA * TIME
PI = 3.14159265359D0
C
C LOOP OVER THE FOUR BLADES:
C
DO 10 I=1,4
C
XANG = COS(ANGLE+(I-1)*PI/2.0D0)
YANG = SIN(ANGLE+(I-1)*PI/2.0D0)
DIST = SQRT( (XV-XANG)**2 + (YV-YANG)**2 ) * VV
C
C IF THE NODE COINCIDES WITH THE BLADE:
C
IF (DIST .LT. THICK) THEN
C
IFLAG(KDU) = 1
IFLAG(KDV) = 1
C
C FOR VERTICAL PERIODIC FACE
C
IF (ABS(X) .LT. 0.001D0) THEN
VAL(KDV) = - OMEGA * Y

9-16 © Fluent Inc., Dec-98


MOVING-BODY CONSTRAINT CAPABILITY Examples

C
C FOR HORIZONTAL REFERENCE FACE
C
ELSE IF (ABS(Y) .LT. 0.001D0) THEN
VAL(KDV) = - OMEGA * X
C
C FOR NODES IN CONTINUUM
C
ELSE
VAL(KDU) = - OMEGA * Y
VAL(KDV) = OMEGA * X
ENDIF
C
ENDIF
10 CONTINUE
C
RETURN
END

© Fluent Inc., Dec-98 9-17


Examples MOVING-BODY CONSTRAINT CAPABILITY

Table 9.5: Example 2—FIPREP file

FIPREP
PROBLEM(2-D,NONLINEAR,TRANSIENT)
SOLUTION(SEGREGATED=20)
EXECUTION(NEWJOB)
OPTIONS(UPWINDING)
/
PRINTOUT(NONE)
TIMEINTEGRATION(BACKWARD,NSTEP=10,DT=1.87535378079E-
1,FIXED,NOPRED)
/
DENSITY(CONSTANT=1000)
VISCOSITY(CONSTANT=0.02)
/
ENTITY(TYPE="INNER",FLUID,MOVING)
ENTITY(TYPE="OUTER",FLUID)
ENTITY(TYPE="BAFFLE",PLOT)
ENTITY(TYPE="WALL",PLOT)
ENTITY(TYPE="REFER",SLIP)
ENTITY(TYPE="PERIOD",SLIP)
/
BCPERIODIC(UN,UT,ENTITY,REFERENCE="REFER",PERIODIC="PERIOD",
R1NODE=184,R2NODE=185,P1NODE=191,P2NODE=192,EXCLUSIVE)
BCNODE(VELOCITY,ZERO,NODE=183)
BCNODE(VELOCITY,ZERO,ENTITY="WALL")
BCNODE(VELOCITY,ZERO,ENTITY="BAFFLE")
/
END
CREATE
END

9-18 © Fluent Inc., Dec-98


MOVING-BODY CONSTRAINT CAPABILITY Examples

Figure 9-8: Example 2—Streamline contour plot, Time = 1.31

Figure 9-9: Example 2—Streamline contour plot, Time = 1.50

© Fluent Inc., Dec-98 9-19


Examples MOVING-BODY CONSTRAINT CAPABILITY

Figure 9-10: Example 2—Streamline Contour Plot, Time = 1.69

9-20 © Fluent Inc., Dec-98


MOVING-BODY CONSTRAINT CAPABILITY Examples

9.3.3 3-D Stirred Tank

Consider the three-dimensional model of the stirred tank shown schematically in Figure 9-
11. The tank is 2 meters in diameter, 2.5 meters high, and the bottom of the tank is
rounded. Inside the tank are 4 baffles and 8 rotating blades, arranged in 2 sets of 4 blades
each. Each blade measures 50 by 20 centimeters, and the blades rotate at a speed of 5.235
rads/s or roughly 50 rpm. The tank is filled with water.
The Reynolds number based upon the blade tip speed is 106. Turbulence is modeled by
the automatic mixing length option specified through the VISCOSITY command. The
sides and bottom of the tank are WALL elements. The top of the tank is a slip surface,
and there is no normal velocity through this surface.
The mesh for Example 3 (see Figure 9-12) is composed of 4 continuum entities. The con-
tinuum entity named “blades” represents the sweep of the 8 blades. The group of elements
representing the rotating shaft are not a part of this entity. Only the entity “blades” is con-
sidered as MOVING on the ENTITY command. Again, at each time step, the blades
coincide with nodal points within the mesh; therefore, the mesh has a fan-type topology
within the swept region. A fixed time step is used, in conjunction with the mesh topology,
to ensure that the blade progresses correctly from node to node.
Table 9.6 shows the user subroutine, USRMBC, for Example 3. The structure of the
USRMBC subroutine is similar to that of the previous two examples. Again, the elements
representing the rotating shaft are not considered as moving. All nodes within the moving
entity “blades” are thus potential candidates for the moving-body constraint. The axis of
rotation ( x = 0, y = 0 ) is already excluded from the influence of the subroutine.

Table 9.7 shows the FIPREP file for Example 3. Since the mesh for this example was
originally constructed in centimeters, the SCALE command is included to convert from
centimeters to meters. The segregated solution method is employed, due to the size of the
model.
Results for Example 3 are shown in Figure 9-13 and Figure 9-14, and the FIPOST file to
generate the results is given in Table 9.8. The blades are visualized in three-dimensions as
an iso-surface of the value 0.9 of SPECIES = 14.

© Fluent Inc., Dec-98 9-21


Examples MOVING-BODY CONSTRAINT CAPABILITY

blades
baffles

Figure 9-11: Example 3—Three-dimensional stirred tank

Figure 9-12: Example 3—Element mesh plot

9-22 © Fluent Inc., Dec-98


MOVING-BODY CONSTRAINT CAPABILITY Examples

Table 9.6: Example 3—USRMBC subroutine

SUBROUTINE USRMBC (VAL, NODE, TIME, SOL, ID, NDOF, NUMNP,


1 LDOFU, CONSTR, NODEPR, XYZ, IFLAG, IERR)
C
#include "IMPLCT.COM"
#include "PARUSR.COM"
DIMENSION SOL(*),ID(NUMNP,NDOF),LDOFU(*),XYZ(NUMNP,*)
DIMENSION IFLAG(*),VAL(*),NODEPR(*),CONSTR(*)
C
ZRO = 0.D0
C
C RETRIEVE NODAL COORDINATES:
X = XYZ(NODE,1)
Y = XYZ(NODE,2)
C
C INITIAL CALCULATIONS:
VV = SQRT(X**2+Y**2)
XV = X / VV
YV = Y / VV
C
THICK = 0.015D0
OMEGA = 5.235D0
ANGLE = OMEGA * TIME
PI = 3.14159265359D0
C
C LOOP OVER THE FOUR BLADES:
C
DO 10 I=1,4
C
XANG = COS(ANGLE+FLOAT(I-1)*PI/2.0D0)
YANG = SIN(ANGLE+FLOAT(I-1)*PI/2.0D0)
DIST = SQRT( (XV-XANG)**2 + (YV-YANG)**2 ) * VV
C
C IF THE NODE COINCIDES WITH THE BLADE:
C
IF (DIST .LT. THICK) THEN
C
IFLAG(KDU) = 1
IFLAG(KDV) = 1
IFLAG(KDW) = 1
C
VAL(KDU) = - OMEGA * Y
VAL(KDV) = OMEGA * X
ENDIF
10 CONTINUE
RETURN
END

© Fluent Inc., Dec-98 9-23


Examples MOVING-BODY CONSTRAINT CAPABILITY

Table 9.7: Example 3—FIPREP file

FIPREP
$w = 5.235
$dt = 7.50141512319e-2
PROBLEM(3-D,NONLINEAR,TRANSIENT,TURB)
SCALE(VALUE=0.1)
EXECUTION(NEWJOB)
SOLUTION(SEGREGATED=40,CR=500,CGS=500,PRECON=21)
/
PRESS(MIXED,DISCONTINUOUS)
/
PRINTOUT(NONE)
POST(NBLOCK=1)
5,25,5
TIMEINTEGRATION(BACKWARD,NSTEP=25,DT=$DT,FIXED,NOPRED)
/
DENSITY(CONSTANT=1000)
VISCOSITY(CONSTANT=0.001,MIXLENGTH)
/
ENTITY(TYPE="BLADES",FLUID,MOVING)
ENTITY(TYPE="INNER",FLUID)
ENTITY(TYPE="MIDDLE",FLUID)
ENTITY(TYPE="OUTER",FLUID)
/
ENTITY(NAME="TOP",PLOT)
ENTITY(NAME="BAFFLE",PLOT)
ENTITY(NAME="WALL",WALL)
/
BCNODE(VELOCITY,ZERO,ENTITY="WALL")
BCNODE(VELOCITY,ZERO,ENTITY="BAFFLE")
BCNODE(UZ,ZERO,ENTITY="TOP")
/
END

9-24 © Fluent Inc., Dec-98


MOVING-BODY CONSTRAINT CAPABILITY Examples

Figure 9-13: Example 3—Velocity vector plot, Time = 1.88

Figure 9-14: Example 3—Velocity vector plot (side view), Time = 1.88

© Fluent Inc., Dec-98 9-25


Examples MOVING-BODY CONSTRAINT CAPABILITY

Table 9.8: Example 3—FIPOST file

FIPOST
/
TITLE
THREE-DIMENSIONAL STIRRED TANK
/
POOR(OFF)
WINDOW(XVIEW=0.5,YVIEW=1,ZVIEW=1)
/
MESH
/
SUPER(ON)
TIME(STEP=25)
HEADING(OFF-SCALED)
PLANE(SECTION,NOBODY,ENTITY="BAFFLE")
EDGE
PLANE(SECTION,NOBODY,FUNCTION,SPECIES=14)
0.9
MESH
PLANE(SECTION,NOBODY,PVECTOR)
0 0 0 0 1 0
HEADING(ON)
VECTOR
SUPERIMPOSE(OFF)
/
WINDOW(YVIEW=1)
VECTOR(FACTOR=100)
/
END

9-26 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Overview

10. NONDIMENSIONALIZATION
10.1 Overview
The equations that FIDAP solves are listed and explained in the FIDAP Theory Manual.
These equations are expressed in terms of physical variables, such as density, velocity,
temperature, etc. When being solved in this form the equations require that data should be
input in dimensional (that is, physical) units. FIDAP accepts any system of units (S.I.,
British, etc.) and outputs the results in the same system of units. For example, if velocity
boundary conditions are applied in cm/sec the velocities output by the program should be
understood to be also in cm/sec. Although it does not matter what system of units is
chosen, it is important that the units must be self-consistent—that is, that all data input
quantities must be in the same system of units; otherwise the results will not be mean-
ingful.
FIDAP also has the capacity of solving the governing equations in dimensionless form. In
this case the input data are dimensionless parameters, such as Reynolds number, and the
output results are then dimensionless quantities.
Dimensionless formulation of problems has many significant advantages. Scaling the fun-
damental variables with respect to typical values and constructing dimensionless parame-
ters provides a measure of the relative importance of the various terms in the equations
and identifies the dominant physical phenomena. This in turn allows an estimate of the
difficulty of the problem and gives insight into meshing requirements as, for example,
when thin boundary layers occur. Nondimensional formulation can sometimes result in a
reduction of the large differences in orders of magnitude that can occur between terms in
an equation. It also facilities the performance of parametric studies through the variation
of a single parameter.
In this chapter the procedures to be followed when setting up a problem in dimensionless
form are outlined and the most important dimensionless parameters that appear in CFD
simulations indicated.
Nondimensionalization means, in essence, choosing a typical or characteristic value for
each variable and relating the actual value of the variable to the characteristic value. In
virtually any kind of flow, two crucial variables are the velocity vector and the position
vector. These quantities, ui and xi , respectively, are scaled with respect to a characteristic
speed U and a characteristic length L, and dimensionless velocities and distances, ui* and
xi* , are defined by the relations

ui* = ui / U , xi* = xi / L . (10.1.1)

(Here and subsequently, * denotes a dimensionless quantity.)

© Fluent Inc., Dec-98 10-1


Overview NONDIMENSIONALIZATION

It should be noted that in all but the simplest problems there is more than one length L or
speed U available for scaling. You are free to choose which pair you wish to use, and the
results of your simulations may not depend very much on your choice. But in the interests
of consistency it is definitely advisable to choose L and U to be at the same geometric
location.
For example, if U is a measure of the velocity at the inflow to your computational do-
main, L should be a measure of the size of that inflow region; L should not be a length
taken from the outflow of the computational domain.
We turn now to a more detailed examination of various types of flows and look at the
nondimensionalizations involved in each case.

10-2 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Isothermal Flows

10.2 Isothermal Flows


The simplest type of flow is laminar, isothermal flow with constant density and viscosity.
The equations of conservation of mass and momentum in this case are
ui,i = 0 (10.2.1)

and
FG ¶u + u u IJ = - p + t
r
H ¶t K
i
j i, j ,i ij , j . (10.2.2)

Applying the change of variables (10.1.1) to equation (10.2.1) we see that the latter is in-
variant when nondimensionalized. In equation (10.2.2) we introduce the transformation
(10.1.1) and, in addition, the scalings
t* = tU / L , p* = p / rU 2 , t *ij = t ij L / mU a f . (10.2.3)

Equation (10.2.3) says that the time scale is L/U, the pressure scale is rU 2 and the scale
for the shear stress is mL / U The first of these is a natural definition of time, the second
comes from well-known physical considerations, such as Bernoulli’s equation, and the
third derives from the fact that
t ij = mdij (10.2.4a)

where dij is the shear rate tensor with dimension U/L. The dimensionless form of
(10.2.4a) is therefore
t *ij = dij* . (10.2.4b)

Substituting (10.1.1) and (10.2.3) into equation (10.2.2), and making some rearrange-
ments, the latter becomes

¶ui* 1 *
+ u*j ui*, j = - p,*i + dij , j (10.2.5a)
¶t* Re
where
rUL
Re = (10.2.6)
m
is the Reynolds number. This is a dimensionless parameter, probably the most important
one in all fluid dynamics. It measures the relative importance of inertial forces to viscous
forces. When Re is large, inertial forces dominate; boundary layers form and the flow
may become turbulent. When Re is very small, the flow is slow and viscous forces
dominate.

© Fluent Inc., Dec-98 10-3


Isothermal Flows NONDIMENSIONALIZATION

The foregoing procedure has established that the characteristics of an isothermal flow of a
fluid with constant density and viscosity are completely determined by a single parameter,
the Reynolds number. However, there is not a unique representation of the momentum
equation. A simple rearrangement of (10.2.5a) results in

FG ¶u
*
IJ
H ¶t K
+ u*j ui*, j = - p,**i + dij* , j
i
Re * (10.2.5b)

where
p** = Re p* . (10.2.7)

Equations (10.2.5a) and (10.2.5b) are identical except for the different interpretations of
pressure. Observe that, by contrast with the scaling of pressure in (10.2.3), we now have
p** = pL / mU . (10.2.8)

Either (10.2.5a) or (10.2.5b) can be used. Experience shows that (10.2.5a) is a good
choice for high Reynolds number flows, while (10.2.5b) is a good choice for low
Reynolds number flows. If (10.2.5a) is chosen, the input to FIDAP requires that the
density (that is, the coefficient of the term on the left-hand side of the equation) be set
equal to 1, and the viscosity be set equal to 1/Re. If (10.2.5b) is used, the density in
FIDAP should be set equal to Re and the viscosity to 1.
If you decide to nondimensionalize your problem, you must remember also to nondimen-
sionalize the boundary conditions. In the case of velocity boundary conditions, all pre-
scribed velocities should be scaled with respect to the reference velocity U. In the case of
flux boundary conditions, however, a little care needs to be taken. Flux boundary condi-
tions are prescribed values for various components of the (total) stress s ij , defined by

s ij = - pd ij + t ij . (10.2.9)

At this point, you have two options for the nondimensionalization of the stress. You may
choose
s ij = s *ij rU 2 (10.2.10a)

in which case (10.2.9) becomes

a
s *ij = - p*d ij + 1 / Re t *ij f (10.2.11a)

or you may choose


s ij = s ** a
ij mU / L f (10.2.10b)

10-4 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Isothermal Flows

in which case (10.2.9) becomes

ij = - p d ij + t ij
s ** ** *
. (10.2.11b)

Either of these is a viable option. However, what is critically important is that if you
choose (10.2.10a), then you must use the form (10.2.5a) of the momentum equation while
if you choose (10.2.10b), you must use (10.2.5b).
This highlights one of the most important things to remember about nondimensionaliza-
tion: you must have the same scalings for the boundary conditions as you do for the gov-
erning equations. If you do not do this, you will not be solving the problem you want to
solve.

10.2.1 Gravitational Body Force

In most isothermal flow problems it is not necessary to include a gravitational body force
term in the governing equation, since this effect is balanced by a hydrostatic pressure gra-
dient. However, in problems with free surfaces or fluid-fluid interfaces it may be neces-
sary to include this term explicitly. In this case equation (10.2.2) becomes

FG ¶u + u u IJ = - p + rg + t
r
H ¶t K
i
j i, j ,i i ij , j . (10.2.12)

Assuming that the vector gi is constant in magnitude and direction, we can use the same
scalings as before, and write
gi = ggi* (10.2.13)

where g is the magnitude of gravity and gi* is now a unit vector in the direction of gravity.
Then we obtain

¶ui* 1 *
+ u*j ui*, j = - p,*i + Fr -1 gi* + t ij , j (10.2.14a)
¶t* Re
where the dimensionless parameter Fr, defined by

Fr = U 2 gL (10.2.15)

is called the Froude number. It is a measure of the relative importance of inertial forces to
gravitational forces. When Fr is large gravitational forces are unimportant, and con-
versely. The alternative form to (10.2.14a) is

FG ¶u *
IJ
H ¶t K
+ u*j ui*, j = - p,**i + Re Fr -1 gi* + t *ij , j
i
Re *
. (10.2.14b)

© Fluent Inc., Dec-98 10-5


Isothermal Flows NONDIMENSIONALIZATION

In FIDAP the magnitude of gravity should be set equal to Fr -1 in both cases. The mag-
nitude of gravity is entered using the GRAVITY command.

10.2.2 Variable Viscosity

In the preceding discussion, it has been assumed that the density and viscosity of the fluid
are constant. The case of variable density will be discussed in more detail below; we con-
tinue to assume for the present that it is constant. But a few comments about what to do
when the viscosity varies are now appropriate.
The Reynolds number, which involves viscosity, is by definition a constant (as are all
dimensionless parameters). If in a particular problem the viscosity is varying, the defini-
tion (10.2.6) of the Reynolds number must be modified to read
rUL
Re = (10.2.16)
m0
where m 0 is a reference viscosity. Likewise, in the scalings defined above, such as
(10.2.3), (10.2.4), (10.2.8), and (10.2.10b), m must be replaced by m 0 .

The choice of m 0 is at your discretion. If the viscosity is a function of shear rate, a good
choice of m 0 is often the zero shear rate viscosity. If the viscosity varies with temperature,
m 0 should be the value at some specific reference temperature. Clearly there could be
more than one option and there are no strict rules, only a guideline suggesting that you
select a “physically reasonable” value for m 0 .

Once a reference value m 0 has been chosen, you need to nondimensionalize the viscosity
by writing
m = m0m* . (10.2.17)

The quantity m * is a dimensionless viscosity; obviously m * = 1 when the true (that is,
physical) viscosity has the value m 0 . From (10.2.17) and the scalings

FG m U IJ t FG U IJ d
t ij =
H LK , dij =
H LK
0 * *
ij ij

we now obtain, in place of (10.2.4a), the dimensionless relation


t *ij = m * dij* . (10.2.18)

This means that equations (10.2.5a) and (10.2.5b) respectively become

10-6 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Isothermal Flows

¶ui*
¶t *
+ u*j ui*, j = - p,*i +
1 * *
Re
c
m dij h ,j (10.2.19a)

and

FG ¶u
*
IJ c h
H ¶t K
+ u*j ui*, j = - p,**i + m * dij*
i
Re * ,j . (10.2.19b)

In case (a) the FIDAP input for viscosity is through a curve or subroutine that defines the
function; m * /Re in case (b) the values set are for the function m * .

An important special case belonging to this category is that of the non-Newtonian fluid.
This is characterized by the fact that the viscosity m is dependent on the shear rate mag-
nitude D defined by
D 2 = 21 dij dij (10.2.20)

(summation over i and j being assumed). The scale for D is evidently U/L, so it may be
nondimensionalized by setting

a f
D = U L D* . (10.2.21)

Thus we can write for the viscosity

a f
m D = m 0 m * D* c h . (10.2.22)

We shall illustrate how to implement these nondimensionalizations by considering a few


special cases.

Bingham Fluid

As noted in the FIDAP Theory Manual, Chapter 2, the viscosity for a Bingham fluid in
dimensional form is
g (10.2.23)
m = m0 + , if t ³ g
D
where g is yield stress, t is the magnitude of the stress tensor,
t 2 = 21 t ijt ij ,
and m 0 is a constant. When t < g the viscosity is effectively infinite and the material
does not move. Using m 0 as the reference value of viscosity, and noting that

© Fluent Inc., Dec-98 10-7


Isothermal Flows NONDIMENSIONALIZATION

b g
g = m 0U L g * , t = m 0U L t * b g (10.2.24)

we see that the dimensionless equivalent of (10.2.23) is

g* (10.2.25)
m* = 1 + , if t * ³ g * .
D*
FIDAP requires that two numbers be entered as data records associated with the VIS-
COSITY(BINGHAM) command, namely the values of m 0 and g if the dimensional form
is being used. With nondimensionalization the values entered depends on whether you are
using format (a) or format (b) of the momentum equation. Most flows involving non-
Newtonian fluids are low Reynolds number flows; therefore we recommend that you use
the (b) format. If you do—that is, if you are using equation (10.2.19b), the appropriate
entries are
viscosity = 1 , yield stress = g * . (10.2.26b)

But if you use (10.2.19a) your entries should be

viscosity = 1/Re , yield stress = g * Re (10.2.26a)

Power-Law Fluid

FIDAP expresses the viscosity in this case as

m=
RS m KD
0
n- 1
, D ³ D0
Tm KD n- 1
. (10.2.27)
0 0 , D < D0

Please note that in this formulation it is assumed that m 0 has the dimensions of viscosity;
this implies that the constant K must have dimensions D01- n . Because of this we nondi-
mensionalize it by writing

K= U L a f 1- n
K* (10.2.28)

and then using (10.2.17) and (10.2.21) we obtain the following dimensionless equivalent
of (10.2.27):

m* =
RS K D * *n - 1
, D* ³ D0*
TK D
, (10.2.29)
* *n - 1
0 , D* < D0*

The FIDAP data record for VISCOSITY(POWERLAW) requires that four values be pre-
scribed, namely m 0 , K, n, and D0 . If we use (10.2.19b), our settings would be

10-8 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Isothermal Flows

1 K* n D0* (10.2.30b)

while if we use (10.2.19a), our settings would be


1 Re K* n D0* . (10.2.30a)

Carreau Fluid

This has the viscosity-shear rate relation

b
m = m ¥ + m 0 - m ¥ 1 + K 2 D2 gc h a n - 1f / 2

(10.2.31)

where K is a constant (of dimension L/U), and m 0 , m ¥ are respectively zero and infinite
shear rate viscosities. Either of these could be, used as the reference viscosity. If we use
m 0 , then (10.2.31) becomes

m* =
m¥ FG
m IJ c h a n - 1f / 2

H
+ 1 - ¥ 1 + K * 2 D* 2
K
(10.2.32)
m0 m0
The data record entries for VISCOSITY(CARREAU) are, in order, zero shear rate vis-
cosity, K, n, infinite shear rate viscosity. With (10.2.19b) these become
1 K* n m¥ m0 (10.2.33a)

while with (10.2.19a) they become


1 Re K* n m ¥ m 0 Re . (10.2.33a)

As indicated above, we could have used m ¥ as the reference viscosity. In that case we
would obtain, in place of (10.2.32), the expression

FG m IJ c h a n - 1f / 2
m* = 1 +
Hm K
- 1 1 + K * 2 D* 2
0 (10.2.34)
¥

with settings
m0 m¥ K* n 1 (10.2.35a)

in place of (10.2.33b) and


m 0 m ¥ Re K* n 1 Re (10.2.35a)

in place of (10.2.33a). If you select m ¥ as your viscosity scale, however, you must
remember to use the same value in your definition of Reynolds number.

© Fluent Inc., Dec-98 10-9


Energy Equation NONDIMENSIONALIZATION

10.3 Energy Equation


The equation of energy conservation for an incompressible fluid has the form

FG ¶T + u T IJ = ckT h
rc p
H ¶t j
K
,j ,j ,j +H (10.3.1)

where c p is specific heat at constant pressure, k is conductivity, T is temperature and H is


a combination of heat generation and/or dissipation terms. As indicated in the FIDAP
Theory Manual, Chapter 2, H may include applied sources and sinks, viscous dissipation
and other contributions.
Except in the almost trivial case where only a conduction problem (ui = 0) is being
solved, equation (10.3.1) is coupled with the equations of mass and momentum conser-
vation through the velocity field ui . Also, the density r appears in (10.2.2) as well as in
(10.3.1). For this reason, it is essential that scaling factors used to nondimensionalize
(10.3.1) should be consistent with those used to nondimensionalize (10.2.2).
To begin with, let us suppose that all the physical quantities r , c p , k, and m are constant.
The most important dimensionless parameter for the energy equation is the Prandtl
number Pr, defined by
mc p
Pr = . (10.3.2)
k
This quantity measures the relative magnitudes of molecular diffusion to thermal diffu-
sion. Another useful parameter is the (thermal) Peclet number, Pe T , defined by
Pe T = Pr Re (10.3.3)

where, Re is the Reynolds number given by equation (10.2.6). By combining (10.2.6) and
(10.3.2), we see that
rc pUL
Pe T = (10.3.4)
k
which represents a ratio of the relative importance of advection to diffusion. Note that,
because of (10.3.3), only two of the three parameters Re, Pr, Pe T are independent.

To nondimensionalize (10.3.1) we introduce the scales L and U as before, and the same
transformations for t and ui . For the temperature is it usually convenient to write

T = T0 + DT T * (10.3.5)

10-10 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Energy Equation

where T0 is a reference or ambient value of the temperature, and DT is some typical tem-
perature difference for the problem under consideration (for example, DT could be the
difference between the anticipated values of maximum and minimum temperature). In
some flow problems where the temperature perturbations are due to a prescribed bound-
ary flux q or a prescribed source/sink H, it may not be possible to set DT from antici-
pated extremum temperature values. In these cases, DT may be set using one of the
following expressions:
qL H L2
DT = or DT = .
k k
As was the case with the momentum equation, there is more than one way to write the
dimensionless form of (10.3.1). One useful option is to define a dimensionless source
term H * by

F L I H,
H* = GH rc UDT JK
p
(10.3.6a)

whereupon we obtain

¶T * 1
+ u*j T, *j = T, *jj + H * . (10.3.7a)
¶t * Pe T

Another option is to define

FG L IJ H
2
H ** =
H kDT K (10.3.6a)

which leads to

FG ¶T *
IJ
Re Pr
H ¶t*
K
+ u*j T, *j = T, *jj + H ** . (10.3.7a))

Notice the parallels between the options labeled (a) here and in the previous section; and
likewise between those labeled (b). Of course this was done deliberately and has the
effect of making the FIDAP input self-consistent. FIDAP requires that the values of
density, viscosity, specific heat and conductivity be specified. If option (a) is chosen
throughout, these entries will be respectively
1 1 Re 1 1 Pe T (10.3.8a)

while, if option (b) is chosen, they will be


Re 1 Pr 1 . (10.3.8b)

© Fluent Inc., Dec-98 10-11


Energy Equation NONDIMENSIONALIZATION

(See Table 10.1, below, for a summary of these settings).


When the term H represents or includes a contribution from applied sources and sinks,
H = qs , (10.3.9)

the scalings (10.3.6a) or (10.3.6b) are applied directly to qs , yielding immediately dimen-
sionless quantities qs* or qs** . When H represents or includes a contribution from viscous
(shear) heating, a little care must be exercised. The viscous dissipation is defined by
H = mD 2 (10.3.10)

where D 2 is defined by equation (10.2.20). Using the scalings (10.2.21) and (10.3.6a),
and introducing a new parameter Br, the Brinkman number, defined by

mU 2 (10.3.11)
Br =
kDT
(10.3.10) transforms to

a
H * = Br Pr f Re1 D *2
. (10.3.12a)

The alternative (b) form is


H ** = Br D*2 . (10.3.12b)

The appropriate input data for viscous dissipation is entered using the OPTIONS
(DISSIPATION) command (see Table 10.1).

10.3.1 Boundary Conditions

As is the case with the momentum equation, the boundary conditions must be nondimen-
sionalized in a consistent manner. The possible boundary conditions for the energy
equation are described in Chapter 2 of the FIDAP Theory Manual. One possibility is that
the temperature is prescribed on a segment of the boundary. In this case the prescribed
temperature should be nondimensionalized according to the expression (10.3.5). A second
possibility is that the heat flux is prescribed on a segment of the boundary:

c h
q = - kT, j n j = qa (10.3.13)

where qa is the value of the applied heat flux and n j is the unit normal to the surface. The
correct nondimensionalization of (10.3.13) depends on whether you are using format (a)
or (b) of equation (10.3.7). If you are using (b), then you should use the settings

10-12 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Energy Equation

FG kDT IJ q FG kDT IJ q
H LK H LK
q= *
, qa = * (10.3.14b)
a

so that (10.3.13) becomes


q* = -T, *j n j = qa* . (10.3.15b)

The value that you input using the BCFLUX command is then qa* as defined by

FG kDTPe IJ q FG kDTPe IJ q
H L K H L K
q= T *
, qa = T * (10.3.14a)
a

which gives
1
q* = - T, *j n j = qa* . (10.3.15a)
Pe T

The value input with the BCFLUX command is the qa* defined by (10.3.14a).

The third possibility is that a convective heat transfer coefficient hc is defined, and the
boundary condition is

c h
q = - kT, j n j = hc T - Tc b g (10.3.16)

where Tc is a reference temperature for convective heat transfer. Tc may or may not be the
same as T0 . Again, the correct scaling depends on whether format (a) or format (b) is
being used. In the case of format (b) you should set
Lhc Tc - T0
hc* = , Tc* = (10.3.17b)
k DT
so that (10.2.16) becomes

c
q* = -T, *j n j = hc* T * - Tc* h . (10.3.18b)

The dimensionless heat transfer coefficient hc* and the dimensionless reference tempera-
ture Tc* , defined by (10.3.17b), are the values to be entered in the input file to FIDAP
using the HTRANSFER command. If format (a) is being used, then you need to set
Lhc T - T0
h*c = , Tc* = c (10.3.17a)
kPe T DT
leading to

© Fluent Inc., Dec-98 10-13


Energy Equation NONDIMENSIONALIZATION

q* = -
1
Pe T
c
T, *j n j = hc* T * - Tc* h . (10.3.18a)

The value of the dimensionless heat transfer coefficient to be entered in the input file is
now that defined by (10.3.17a).
The fourth possibility is that a radiative heat transfer coefficient hr is defined, and the
boundary condition is

c h b g c
q = - kT, j n j = hr T - Tr = eS T 4 - Tr4 h (10.3.19)

where e is the emissivity and S is the Stefan-Boltzmann constant. Tr is a reference tem-


perature, which may or may not be the same as T0 . Using the same procedure as before
we introduce the dimensionless quantities

DT 3 LS Tr T (10.3.20b)
S* = , T r* = , T 0* = 0
k DT DT
leading to

c
q* = -T , j*n j = eS* T * + T 0* h 4
- Tr*4 (10.3.21b)

or

DT 3 LS Tr T0 (10.3.20a)
S* = , T r* = , T 0* =
kPe T DT DT
which leads to

q* = -
1
Pe T
c
T , j*n j = eS* T * + T 0* h 4
- Tr*4 . (10.3.21a)

The appropriate value of the dimensionless Stefan-Boltzmann, together with the values of
T 0* and T r* as defined above, need to be entered as FIDAP input data using the EMIS-
SIVITY command.

10.3.2 Variable Properties

As was the case for the momentum equation, some modifications to the preceding discus-
sion need to be made if the material properties are not constant. Suppose for example that
both specific heat and conductivity are variable while density and viscosity are constant.
We need to choose appropriate reference values of these quantities, denoted c p0 and k0 ,
and to introduce the scalings

10-14 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Energy Equation

c p = c p0 c p* , k = k0 k* . (10.3.22)

Then equation (10.3.7a) now takes the form

FG ¶T *
IJ k* *
c*p
H ¶t
*
K
+ u*j T, *j =
Pe T
T, jj + H * (10.3.23a)

while (10.3.7b) becomes

FG ¶T *
IJ c h
Re Pr c*p
H ¶t K
+ u*j T, *j = k*T, *j + H ** . (10.3.23b)
* ,j

The dimensionless parameters appearing here are now defined as


mc p 0 rc p0UL
Pr = , Pe T = (10.3.24)
k0 k0

while the nondimensionlizations (10.3.6) of the source term H need to be similarly modi-
fied to use the reference values k0 and c p0 .

The FIDAP data entries for specific heat and conductivity will, if format (a) is used, be
curves or subroutines for the functions c*p and k* Pe T , respectively. If format (b) is used,
they will be curves or subroutines for the functions Pr c*p and k* , respectively.

10.3.3 Conjugate Heat Transfer

All the discussion until now has been restricted to the case where the computational
domain is occupied by a single material. There are, however, frequently problems where
two or more materials, having different properties, are present in a problem. Nondimen-
sionalization in such cases should be performed with some care to ensure consistency.
The guiding principle to remember is that the same scalings should be used in all the
component materials.
A particular, but important, example of this is the conjugate heat transfer problem. Here,
typically, the mass, momentum and energy conservation equations are to be solved in a
fluid region while the energy equation is to be solved in a surrounding or adjoining solid
region. Thus we suppose that equations (10.2.1), (10.2.2) and (10.3.1), with H= 0 for sim-
plicity, apply in the fluid region, with the equation

r s c sp
¶T
¶t
c
= ksT, j h ,j
(10.3.25)

in the solid region, where the superscript s refers to the properties of the solid.

© Fluent Inc., Dec-98 10-15


Energy Equation NONDIMENSIONALIZATION

You should use the same scaling factors as before; in particular (10.3.5) should be used in
both materials to scale the temperature. In the fluid region, then, you will obviously obtain
the same equations as before, namely (10.2.5a) and (10.3.7a) or (10.2.5b) and (10.3.7b),
depending on which format you use. But turning to the equation (10.3.25) in the solid,
you will soon find that there are quite a few different ways to write this equation in
dimensionless form. Which are correct, or is the choice completely open? The answer is
that not all possibilities are correct; your choice must be such that the ratio of conductivi-
ties (ks k ) is preserved after the nondimensionalization. The reason for this is that at the
interface between the fluid and the solid (or indeed, between any two adjoining materials)
two conditions are applied: continuity of temperature and continuity of heat flux. (An
exception to this global rule occurs when there is a gap between the materials, but we do
not consider this case here.) If you nondimensionalize temperature according to (10.3.5),
you will be guaranteed continuity of temperature, because
Tfluid = Tsolid Þ Tfluid
*
= Tsolid
*
.

But the heat flux continuity condition


- kT, j n j fluid = - ksT, j n j solid

will only be maintained in the dimensionless form of (10.3.25) if you ensure that the ratio
(ks k ) has been preserved.

If format (a) is chosen, you will soon see that the correct dimensionless form of (10.3.25)
is
r s c sp ¶T * ks FG IJ
1
rc p ¶t *
=
H K
k Pe T
T, *jj . (10.3.26a)

You will have entered 1 / Pe T for the fluid conductivity; now you must enter
c h
ks kPe T as the solid conductivity. The fluid density and specific heat were both entered
as 1 for this format; for the solid density and specific heat you have numerous options:
you can enter r s r for density and c sp c p for specific heat, or any other combination
c
subject only to the condition that their product is the coefficient r s c sp rc p . h
If format (b) is chosen, you will obtain
r s c sp ¶T * ks * FG IJ
rc p
Re Pr * =
¶t k
T, jj
H K . (10.3.26b)

Now you must enter the solid conductivity as ks k . You have very many choices for the
solid density and specific heat. The option is yours, provided only that the product of the
two is correct.

10-16 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Energy Equation

Note finally, that the density and specific heat of the solid region are only relevant when a
TRANSIENT simulation is being performed. In a STEADY simulation, no explicit values
need to be assigned for r s and c ps in which case default values of 1.0 are automatically
assigned.

© Fluent Inc., Dec-98 10-17


Species Equations NONDIMENSIONALIZATION

10.4 Species Equations


The conservation equation for a species being transported in the fluid is

FG ¶c + u c IJ = cra c h
r
H ¶t K
j ,j c ,j ,j + qc + R (10.4.1)

where a c is the mass diffusivity, qc is a general source term, and R is a general reaction
term. There is an equation of the form (10.4.1) for each species; it is enough, however, to
consider just one equation since all are nondimensionalized in the same way.
We note, first of all, that c is a mass fraction; in other words it is already a dimensionless
quantity and it is unnecessary to apply scaling to it. All other quantities are scaled as
before. The dimensionless parameters that appear are the (mass transport) Peclet number
UL
Pe c = (10.4.2)
ac
and the Schmidt number
m
Sc = . (10.4.3)
ra c
The relationship analogous to (9.2.3) is
Pe c = Sc Re (10.4.4)

As before we obtain two different representations, corresponding to the formats (a) and
(b) respectively. For format (a) we have

FG Lq IJ FG LR IJ
qc* =
H rU K R* =
H rU K
c
, (10.4.5a)

and the equation


¶c * 1
+ u j c, j = c, jj + qc* + R* (10.4.6a)
¶t *
Pe c

and for format (b) we have

FG L q IJ
2
FG L RIJ
2
qc** =
H ra K , R** =
H ra K
c (10.4.5a)
c c

and the equation

10-18 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Species Equations

FG ¶c + u c IJ = c
H ¶t K + qc** + R**
*
Re Sc * j ,j , jj . (10.4.6b)

These, of course, have been derived under the assumption that r and a are constants;
some appropriate modifications need to be made if they are not.
The way in which the relevant values are entered into FIDAP is worth some extra consid-
eration. If you choose format (a), then by comparing (10.4.1) and (10.4.6a) it is clear that
you should set density = 1 and diffusivity = 1 / Pe c . This is perfectly acceptable, since you
have already set density = 1 for this format when you nondimensionalized the momentum
equation. Format (b) is a little more complicated. Comparing (10.4.1) and (9.3.6b) you
might be inclined to set the density equal to ReSc. However, you cannot do this because
previously, in the momentum equation, you set density = Re. FIDAP provides a way of
overcoming this by introducing a multiplier, called the capacity, in each species equation.
The capacity is not a physical quantity; it is simply a device useful in this mode of nondi-
mensionlization. For the present case, then, given that you already have density = Re, you
would set capacity = Sc to bring about the right combination. Furthermore in this format
you should set diffusivity = 1/Re in order that the product of density and diffusivity will
equal 1.
The following table summarizes the settings for the important quantities discussed so far.

Table 10.1: Basic dimensionless settings

Physical Dimensional Dimensionless Input


Quantity Input Format (a) Format (b)
Density r 1 Re

Viscosity m 1/Re 1

Specific heat cp 1 Pr

Conductivity k 1 / Pe T 1

Dissipation 1 Br/Pr Br

Diffusivity ac 1 / Pe c 1/Re

Capacity 1 1 Sc

© Fluent Inc., Dec-98 10-19


Species Equations NONDIMENSIONALIZATION

10.4.1 Boundary Conditions

As was the case with the momentum and energy equations, the boundary conditions must
be nondimensionalized in a consistent manner. The boundary conditions for each species
equation are described in Chapter 2 of the FIDAP Theory Manual. One possibility is that
the value of the species concentration is prescribed on a segment of the boundary. This
condition is already effectively in dimensionless form, since concentration is a dimen-
sionless variable. A second possibility is that the mass flux is prescribed on a segment of
the boundary:

c
qm = - rac, j n j = qm h (10.4.7)

where qm is the value of the applied mass flux. The correct nondimensionalization of
(10.4.7) depends on whether you are using format (a) or (b). If you are using (b) then you
should set

FG ra IJ q FG ra IJ q
H LK H LK
qm = *
, qm = * (10.4.8b)
m m

so that (10.4.7) becomes


qm* = - c, j n j = qm* (10.4.9b)

and then the value you input on the BCFLUX command is the value of qm* defined by
(10.4.8b). If you are using format (a), then the correct scalings are

FG raPe IJ q FG raPe IJ q
H L K H L K
qm = c *
, qm = c * (10.4.8a)
m m

which gives
1
qm* = - c, j n j = qm* . (10.4.9a)
Pe c

The value input with the BCFLUX command is the qm* defined by (10.4.8a).

The third possibility is that a mass transfer coefficient hm is defined, with a nonlinear
mass transfer law of the form

qm = - crac hn ,j j = hm c n . (10.4.10)

If using format (b) you should define

10-20 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Species Equations

FG ra IJ h
hm =
H LK
*
m
(10.4.11b)

with qm* defined as in (10.4.8b). Then (10.4.10) becomes

qm* = - c, j n j = hm* c n . (10.4.12b)

The value input on the SPTRANSFER command is the value of hm* given by (10.4.11b).
If format (a) is being used, you should define

FG ra IJ h
hm =
H LPe K
* (10.4.11a)
m
c

with qm* defined as in (10.4.8a). Then (10.4.10) becomes

1
qm* = - c, j n j = hm* c n . (10.4.12a)
Pe c

The value input on the SPTRANSFER command is the value of hm* given by (10.4.11a).

A fourth possibility is that a chemical reaction is specified on a segment of the boundary.


This condition has the form
- rac, j n j + cr = rs (10.4.13)

and
ru j n j = r (10.4.14)

where rs represents the surface reaction term of the species and r denotes the sum of this
reaction and the reaction for the carrier gas. (Note that (10.4.13) and (10.4.14) apply to
the case where there is only one species; the generalization to the case of several species
will be discussed below.) We can nondimensionalize (10.4.13) and (10.4.14) by writing
rs = ara / Lfr , *
s r= ara / Lfr *
, ui = Uu*i (10.4.15b)

which reduces these equations to


1 *
- c, j n j + cr * = rs* , u *j n j = r . (10.4.16b)
Pe c

This setting corresponds to format (b). Alternatively we can put


rs = rUr *s , r = rUr * , ui = Uu*i (10.4.15a)

which gives

© Fluent Inc., Dec-98 10-21


Species Equations NONDIMENSIONALIZATION

1
- c, j n j + cr * = rs* , u j* n j = r * . (10.4.16a)
Pe c

This corresponds to format (a).

10.4.2 Multiple Species Equations

When more than one species equation is present in a problem, the nondimensionalization
procedure can be performed as for the case of a single species equation, with only a few
obvious modifications. However, some care needs to be taken in writing the dimension-
less formats of the equations to be solved. By way of illustration consider the situation of
N equations of the form (10.4.1):

FG ¶c + u c IJ = ra c
r
H ¶t K + Rn , n = 1,..., N
n
j n, j n n, jj . (10.4.17)

Here we have assumed that the diffusivity a n is different for each species but is constant
for each. The source term has been omitted for simplicity.
Clearly there will be a distinct Schmidt number and a distinct Peclet number for each
species:
m rLU
Sc n = , Pe n = .
ra n an (10.4.18)

Using our standard nondimensionalizations corresponding to format (a),


u*i = ui / U , t* = tU / L , R n* = LRn / rU (10.4.19a)

we obtain
¶cn * 1
+ u j cn, j = cn, jj + R n* . (10.4.20a)
¶t * Pe n

This is directly analogous to (10.4.6a), and the diffusivities of the species are set equal to
1 / Pe n , n=1,...N, in FIDAP. The boundary conditions for this option are also direct gen-
eralizations of those derived above for the format (a) case, where Pe c replaced by Pe n
whenever it occurs.
The generalization of the format (b) procedure can also be effected directly, but in this
case you need to exercise some caution. You can certainly set

ui tU L2 Rn (10.4.19b)
ui* = , t* = , Rn** =
U L ra n

10-22 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Species Equations

and obtain the N equations

FG ¶c IJ
H ¶t K
+ u j* cn, j = cn, jj + Rn**
n (10.4.20b)
Re Sc n *
.

If you do this you would input density = Re, capacity = Sc n (a different value for each
species) and diffusivity = 1 for each species. You can, moreover, generalize in the same
way the boundary conditions (10.4.9b), (10.4.11b) and (10.4.12b) and make the appropri-
ate entries on the BCFLUX or SPTRANSFER commands respectively. But if you need a
chemical reaction boundary condition, this approach will not work. The condition you
have is
- ra n c, j + cn r = rn , n = 1,..., N (10.4.21)

and
ru j n j = r (10.4.22)

subject to the constraint


N +1
r= år
n=1
n . (10.4.23)

You are going to have trouble because each rn has a different scaling factor if you follow
this procedure, namely ra n L , and so it is not clear how to nondimensionalize r and to
satisfy (10.4.23).
The reason for this difficulty is that format (b) does not preserve the relative magnitude of
the mass fluxes. In the original dimensional form the ratio of fluxes for, say, species 1 and
2 is equal to the ratio of their diffusivities a 1 a 2 . In format (b), however, this ratio is 1.

A better way to nondimensionalize multiple species equations, while essentially retaining


the structure of format (b), is to select the diffusivity of one of the species as the reference
value. Say you choose species 1 for this purpose. Then you write
m rLU
Sc 1 = , Pe 1 = (10.4.24)
ra 1 a1
and in place of (10.4.19b) you set

ui tU L2 Rn (10.4.25b)
u*i = , t* = , Rn** = .
U L ra 1
Then you will obtain the equations

© Fluent Inc., Dec-98 10-23


Species Equations NONDIMENSIONALIZATION

FG ¶c IJ FG a IJ c
H ¶t K Ha K
+ u *j cn, j = + Rn**
n n
Re Sc 1 * n , jj . (10.4.26b)
1

The input for this format is density = Re and capacity = Sc 1 for all the species, and dif-
fusivity = a n a 1 (which, of course, reduces to 1 for the reference species).

The chemical reaction boundary conditions (10.4.21)–(10.4.23) are now nondimensional-


ized in a consistent manner. With

FG ra IJ r FG ra IJ r , u = Uu
rn =
H LK , r=
H LK
1 * 1 * *
n i i
(10.4.27b)

they become

b g
- a n / a 1 cn, j n j + cn r * = r n* , u j* n j =
1 *
Pe 1
r (10.4.28b)

and the constraint (10.4.23) is preserved.

10-24 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Buoyancy-Driven Flows

10.5 Buoyancy-Driven Flows


Gradients of temperature and/or concentration in a fluid generate body forces that cause
the fluid to move. This phenomenon is usually called natural convection (or free convec-
tion) and occurs in many different practical applications. When this type of flow occurs in
conjunction with a forced flow, driven, for example, by a pressure gradient, the combina-
tion of the two is known as mixed convection.
The Boussinesq equations are a well-established approximation for describing both natu-
ral and mixed convection. In this model it is assumed that the density is constant (that is,
the problem type is INCOMPRESSIBLE) except for the gravitational force term, in which
density varies with temperature and/or species concentration, thereby generating the natu-
ral convection. The equations are

FG ¶u + u u IJ = - p + mu + br - r g g
r0
H ¶t K
i
j i, j ,i i, jj 0 i (10.5.1)

u j, j = 0 (10.5.2)

r 0 c p FG I
¶T
H ¶t + u T JK = kT j ,j , jj
(10.5.3)

r FG + u c IJ = r a c
¶c
H ¶t K
n (10.5.4)
0 j n, j 0 n n , jj .

These are simplified versions of equations we have discussed previously. Heat sources,
reaction terms, etc., have been omitted for the sake of brevity but they can be reintroduced
if desired. Similarly, material properties have been assumed constant; the modifications
needed if the properties are variable have been indicated earlier. The density r 0 is a con-
stant reference density evaluated at the constant temperature T0 and at the (zero) reference
value of all the species. In FIDAP there are two ways of expressing the density variation
r - r 0 . One is effectively derived from the ideal gas law, and is
L
r - r 0 = - r 0f Mb T bT - T0 g + å b M N / Mn - 1 gOPQ
N n
(10.5.5.I)

where Mn is the molecular weight of the nth species, M N is the molecular weight of the
carrier, and f is a multiplier defined by
1
f=
1+ å
n
b M N / Mn - 1 cn g . (10.5.6)

The other is commonly known as the Boussinesq form, and is

© Fluent Inc., Dec-98 10-25


Buoyancy-Driven Flows NONDIMENSIONALIZATION

LM b g
r - r 0 = - r 0 b T T - T0 + å b c cn
OP
N n
n
Q . (10.5.5.II)

In equations (10.5.5) b T and b cn are the coefficients of volumetric expansion associated


with temperature and with the nth species, respectively.
Note that if there are no species in the problem then f = 1 and the two forms become
identical. The FIDAP designation of the two alternative representations of the buoyancy
force is by the keywords TYP1 and TYP2, respectively, on the DENSITY command.
We now want to consider the best ways of nondimensionalizing the system (10.5.1) –
(10.5.6). As usual, there are several options. The most important thing you need to exam-
ine before doing anything else is: which is more important, forced convection or natural
convection? The answer to this question will determine the right scalings to use. One way
to make this determination is to look at the ratio of the buoyancy force to the inertial
force, namely
br - r g gL
0
(10.5.7)
r 0U 2
where g is the magnitude of gravity and U is the reference velocity for forced flow. If the
ratio (10.5.7) is much less than 1, forced convection dominates; if it is much greater than
1, natural convection dominates; if it is O(1), the two modes of convection are equally
important.
Your scaling factors should be based on this consideration. When the ratio (10.5.7) is
much less than 1, forced flow dominates and so it is logical that you use the same scalings
as in Section 10.2 above. As you do this you will see that once again you have the option
of format (a) or format (b). After a little manipulation you will find that the momentum
equation takes one of the following forms:

¶ui* F FG IJ I
¶t
+ u*j ui*, j = - p,*i +
1 *
Re GH
ui, jj - Fr -1 gi* fb *T T * + f å
nH
MN
Mn
- 1 cn
K JK (10.5.8a.I)

or

¶ui* F I
¶t
+ u*j ui*, j = - p,*i +
1 *
Re GH
ui, jj - Fr -1 gi* b *T T * + å b c cn
n JK (10.5.8a.II)

or

10-26 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Buoyancy-Driven Flows

FG ¶u
*
IJ FG b
+ u*j ui*, j = - p,**i + ui*, jj - Re Fr -1 gi* fb *T T * + f å M N / Mn - 1 cn g IJK
H ¶t K H
i
Re *
n

(10.5.8b.I)
or

FG ¶u
*
IJ F I
Re
H ¶t
i
* GH
+ u*j ui*, j = - p,**i + ui*, jj - Re Fr -1 gi* b *T T * + å b c cn
K n JK . (10.5.8b.II)

Note that here all quantities have been scaled as in the preceding sections, and that the
Reynolds number is defined with respect to the reference density. In all the above equa-
tions we have set
b *T = bDT . (10.5.9)

The magnitude of gravity must be set to Fr -1 for both formulations (10.5.8b.I) and
(10.5.8b.II) above. This is done using the GRAVITY command. In the case under consid-
eration, where forced convection dominates, the nondimensionalizations of the tempera-
ture and species equations are exactly as described in Sections 10.3 and 10.4, so there is
no need to repeat the details here.
Things are quite different if natural convection dominates, or is the only mechanism
present. If you look at the ratio (10.5.7) you will see that the quantity r - r 0 gL r 0 has b g
the dimensions of the square of velocity. When the ratio (10.5.7) is much greater than 1,
or infinite, it seems logical to take as a reference velocity not U, the velocity associated
with the forced flow (especially since this may be zero), but rather the quantity

UN = br - r g gL r
0 0 (10.5.10)

which is a representative velocity for the flow driven by buoyancy forces.


We are going to proceed on the assumption that temperature gradients are what primarily
caused the natural convection. Sometimes this may be false, and the main driving mecha-
nism may be concentration gradients. If so, the discussion that follows will have to be
modified, but the procedure should be reasonably self-evident.
If you look at the two possible forms of r - r 0 as given by (10.5.5.I) and (10.5.5.II),
given that temperature gradients dominate the natural convection, we see that (10.5.10)
can be written in the simpler form

U N = b T DTgL . (10.5.11)

© Fluent Inc., Dec-98 10-27


Buoyancy-Driven Flows NONDIMENSIONALIZATION

(Of course (10.5.10) and (10.5.11) are not equal to each other; we are only trying to
define a scale for the velocity.) This new scaling leads to a new parameter, called the
Grashof number, defined as

r 2 b T DTgL3 (10.5.12)
Gr = .
m2
This is the fundamental parameter for natural convection dominated flows, and in this
sense replaces the Reynolds number. But in reality the change of fundamental parameter
is not much more than a change of name. For if you define the Reynolds number based on
U N , namely

r 0U N L
Re N =
m (10.5.13)

Gr = Re 2N . (10.5.14)

We can now proceed to use the same scalings as in Sections 10.2 – 10.4, bearing in mind
only that the velocity scale is now given by (10.5.11). We also now have new definitions
of the Peclet number, namely

Pe T = Pr Gr , Pe n = Sc n Gr . (10.5.15)

When the Boussinesq form (10.5.5.II) is being used to approximate the buoyancy force, it
is easy enough to show that the dimensionless momentum equations corresponding to
formats (a) and (b) are

¶u*i
+ u *j ui*, j = - p,*i +
1 * LM
ui, jj - gi* T * + å b *c cn
OP
N Q
(10.5.16a.II)
¶t *
Gr n
n

and

FG ¶u
*
IJ
+ u*j ui*, j = - p,**i + ui*, jj - Gr gi* T * + å b *c cn
H ¶t K
Gr i (10.5.16b.II)
* n

where

b *c = b c b *T DT
n n
and p** = Gr p* . (10.5.17)

When the ideal gas approximation (10.5.5.I) is being used a little more care needs to be
taken because the molecular weight expressions cannot be nondimensionalized. After
some rearrangements we find the analogs of the above equations to be

10-28 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Buoyancy-Driven Flows

¶u*i 1 * 1 LM
ui, jj - * gi* fb *T T * + f å
MN FG IJ OP
¶t*
+ u j* ui*, j = - p,*i +
Gr bT N Mn
- 1 cn
H K Q (10.5.16a.I)

and

FG ¶u
*
IJ F FG M IJ I
Gr
H ¶t
i
*
K
+ u*j ui*, j = - p,**i + ui*, jj -
b
Gr
*
T
GH
gi* fb *T T * + f å
HM
N

n K JK
- 1 cn (10.5.16b.I)

where

p** = Gr p* and b *T = bDT . (10.5.18)

The energy and species equations retain the same forms as previously, and it is unneces-
sary to write them down here. The FIDAP settings for the four cases are summarized in
Table 10.2, below.
We note finally that a parameter commonly used in place of the Grashof number is the
Rayleigh number Ra. This is defined by

Ra = GrPr . (10.5.19)

Thus Gr = Ra / Pr may be used in the equations discussed in this section.

© Fluent Inc., Dec-98 10-29


Buoyancy-Driven Flows NONDIMENSIONALIZATION

Table 10.2: Dimensionless settings for natural convection

Physical Dimensional Dimensionless Input


Quantity Input (a)I (a)II (b)I (b)II
Density r0 1 1 Gr Gr

Viscosity m 1 / Gr 1 / Gr 1 1

Gravity g 1 / b T DT 1 1 / b T DT 1

Volume bT b T DT 1 b T DT 1
expansion
(thermal)

Volume bc – b c / b T DT – b c / b T DT
expansion
(species)

Specific heat cp 1 1 Pr Pr

Conductivity k 1 / Pe T 1 / Pe T 1 1

Diffusivity a 1 / Pe c 1 / Pe c 1 / Gr 1 / Gr

Capacity 1 1 1 Sc Sc

10-30 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Compressible Flows

10.6 Compressible Flows


In FIDAP a flow is designated as INCOMPRESSIBLE if the density is constant except
possibly in the gravitational force term, as discussed in the previous section; and it is
designated as COMPRESSIBLE if the density is variable.
FIDAP divides the treatment of compressible flow into two categories, depending on
whether the density is independent of pressure or the density and pressure are related, as,
for example, by the ideal gas law.
If the former case applies, there is very little to add to our previous discussion as regards
nondimensionalization. We have already discussed what to do, for instance, when the vis-
cosity is varying. The same principle applies when the density is varying; you need to
introduce a reference density r 0 and scale density according to the formula

r = r0r* . (10.6.1)

All your dimensionless parameters, such as Reynolds number, should clearly be defined
in terms of the reference density value, and the functional form of r * needs to be input
wherever it appears.
For this case FIDAP restricts you to only two possible forms of density variation, namely

b
r 0 1 - b T T - T0 g
r=
1+ å
FG M IJ (10.6.2)

HM K
- 1 cn
N

n n

or

b
r = r 0 1 - b T T - T0 g Õ d1 - b c i
n
cn n . (10.6.3)

It is evident, then, that the dimensionless forms of the density are, respectively,

1 - b *T T *
r* =
1+ å
FG M IJ (10.6.4)

HM - 1 cn
K
N

n n

or

r * = 1 - b *T T * Õ d1 - b
n
cn nc i (10.6.5)

where b *T = b T DT , as in the previous section.

If natural convection is present FIDAP enters the buoyancy force term as

© Fluent Inc., Dec-98 10-31


Compressible Flows NONDIMENSIONALIZATION

br - r g g 0 i (10.6.6)

which transforms into

c
r 0 g r * - 1 gi* h . (10.6.7)

Following the same procedure as in Section 10.5, you will easily see that dimensionless
forms of the momentum equations for this case are similar to (10.5.16a.I) and
(10.5.16b.I), namely

FG ¶u*
IJ 1 * 1
c h
r*
H ¶t K
+ u*j ui*, j = - p,*i + ui, jj + * gi* r * - 1
i (10.6.8a)
*
Gr bT
and

FG ¶u *
IJ Gr
c h
Gr r *
H ¶t K
+ u*j ui*, j = - p,**i + ui*, jj + gi* r * - 1
i (10.6.8b)
*
b *T
where r * -1 is computed from either (10.6.4) (TYP1 on DENSITY command) or
(10.6.5) (TYP2 on DENSITY COMMAND).

10.6.1 Ideal Gas Law

A very different procedure should be followed to nondimensionalize the equations when


density and pressure are mutually dependent. To illustrate this procedure we shall restrict
ourselves to consideration of the following basic system of equations:

FG ¶u + u u IJ = - p + t
H ¶tr
K
i
j i, j ,i ij , j (10.6.9)

+ c ru h = 0
¶r
(10.6.10)
¶t j ,j

rc FG I
H ¶t + u T JK = kT + t d
¶T
v j ,j , jj
1
2 ij ij - pu j , j (10.6.11)

p = rRT . (10.6.12)

Notice that for these equations the extra stress tensor t ij includes a contribution from the
bulk viscosity— see the FIDAP Theory Manual, Chapter 2.
Before defining the appropriate scalings for the system (10.6.9) – (10.6.12) it must be
emphasized that in compressible/ideal gas problems as posed in FIDAP there is always a
need to set a pressure level. Usually this level, denoted pE , is the pressure at the outflow

10-32 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Compressible Flows

from the domain. Secondly, it should be pointed out that the temperature appearing in the
equation of state (10.6.12) is an absolute temperature.
First of all, let us define some important parameters. The first of these is the ratio of the
specific heat at constant pressure to the specific heat at constant volume:
cp
g = . (10.6.13)
cv

Next we define the quantity U S , the speed of sound at a reference temperature TR , by the
formula

U S = g RTR (10.6.14)

where R is the gas constant. Then, if U is a typical speed of the flow, we define the Mach
number Ma to be
U
Ma = . (10.6.15)
US

The parameters g and Ma, together with the Reynolds and Prandtl numbers, are the fun-
damental parameters in the nondimensionalization of compressible flow problems.
We need to define reference values of density and pressure. We set
pE
rR = (10.6.16)
RTR

in other words we represent the density level in terms of the chosen reference temperature
and the specified pressure level. Then we define
pR = r RU S2 = r Rg R TR . (10.6.17)

Now we scale the quantities in equations (10.6.9) – (10.6.12) as follows:


r = r Rr* , p = pR p* , T = TRT *

ui = U S ui* , b
t = L US t* g (10.6.18)

b g
dij = U S L dij* , t ij = mU S L t *ij b g .

Substituting these into (10.6.9) – (10.6.12) we obtain the following dimensionless system:

© Fluent Inc., Dec-98 10-33


Compressible Flows NONDIMENSIONALIZATION

FG ¶u
*
IJ 1 *
r*
H ¶t K
+ u*j ui*, j = - p,*i +
i
*
t ij , j (10.6.19)
Re S

cr u h
*
¶r
*
+ * *
j ,j =0 (10.6.20)
¶t

FG ¶T *
IJ 1
b 1
g c h
H ¶t K
+ u*j T, *j = T, *jj + BrS Pr dij* - Ec p*u*j , j
-1 * 1 *
g r * 2 t ij (10.6.21)
Pe S Re S

1
p* = r
*
T* (10.6.22)
g

where Re S is the Reynolds number based on the speed of sound:

r RU S L
Re S = . (10.6.23)
m

Pe S and BrS are the corresponding Peclet and Brinkman numbers:


2
mU S
Pe S = Pr Re S , BrS = (10.6.24)
kTR

and Ec is the Eckert number,

U S2
Ec = . (10.6.25)
c pTR

The FIDAP settings for all the parameters in (10.6.19) – (10.6.22), except c p , are pre-
cisely as explained in Section 10.4. The value of the parameter Ec is set with the
REVERSIBLE keyword on the appropriate ENTITY command, while the value of the gas
constant must equal 1 g . In (10.6.21) the coefficient g -1 appears in place of cv , the spe-
cific heat at constant volume. In FIDAP cv is not entered explicitly, but rather it is evalu-
ated from R and c p from

cv = c p - R

or in dimensionless form
1
cv* = c*p - R* = c*p - .
g

Thus in order to ensure that cv* assumes its correct value of g -1 as dictated by (9.5.21). the
value of c*p must be 2 g .

10-34 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Compressible Flows

As before, the scalings (10.6.18) allow different formats of the governing equations, but
the system (10.6.19) – (10.6.22) is the recommended one.
Note finally that the Reynolds number Re based on a characteristic flow velocity is
related to Re S for the formula

Re = Re S Ma . (10.6.26)

© Fluent Inc., Dec-98 10-35


Turbulent Flows NONDIMENSIONALIZATION

10.7 Turbulent Flows


The k - e model introduces two additional equations which have to be nondimensional-
ized in conformity with the equations of momentum, energy, etc. To demonstrate how this
is done let us consider a simple system of equations for turbulent incompressible flow,
including heat transfer, but excluding body forces, sources and species equations. The
governing equations are

FG ¶u + u u IJ = - p + mcu + u h
r0
H ¶t K
i
j i, j ,i i, j j ,i ,j (10.7.1)

F ¶T + u T IJ = clT h
r c G
H ¶t
0 p
K j ,j ,j ,j (10.7.2)

r G
F ¶k + u k IJ = FG m k IJ - r e + m F
H ¶t K Hs K
t
0 j ,j ,j ,j 0 t (10.7.3)
k

F ¶e + u e IJ = FG m e IJ + c ae kfm F - r c e
r G
H ¶t K Hs K
t 2
0 j ,j ,j 1 t 0 2 /k (10.7.4)
e ,j

where
F = 21 dij dij (10.7.5)

is a dissipation function. In these equations m denotes the sum of the laminar and turbu-
lent viscosities,
m = m0 + mt (10.7.6)

where
mt = r 0 cm k2 / e (10.7.7)

l is the total conductivity,

l = l0 + lt (10.7.8)

where
lt = cpm t / s t (10.7.9)

s t being the constant turbulent Prandtl number. The quantities c1 , c2 , cm , s k , s e , s t are


constants.

10-36 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Turbulent Flows

To nondimensionalize the system (10.7.1) – (10.7.4) you can use the same scalings as
before for length, velocity, time, etc., and define dimensionless parameters like the Rey-
nolds number and Prandtl number with respect to the laminar values of viscosity, con-
ductivity, etc. You should also use the scalings

k = U 2 k* , e c
= U3 / L e* h (10.7.10)

for the turbulence quantities, and define


m = m0m* , l = l 0 l* (10.7.11)

where clearly
m
*
= 1 + mt / m0 , *
l = 1+ lt / l0 (10.7.12)

also from (10.7.7) and (10.7.10) we will have


mt = r 0ULm *t (10.7.13)

where
mt
*
= cm k*2 / e * . (10.7.14)

The system (10.7.1) – (10.7.4) now becomes

c h
*
¶ui 1
+ u*j ui*, j = - p,*i + m ui, j + u j ,i
* * *
,j (10.7.15)
¶t Re

c h
*
¶T 1 * *
*
+ u*j T, *j = l T, j ,j (10.7.16)
¶t Pe
¶k
*
FG m k IJ - e + m F
*
+ u*j k,*j =
Hs K
t * * * *
,j ,j t (10.7.17)
¶t k

¶e
*
FG m *
IJ + c ce / k hm F - c e
+ u*j e ,*j =
Hs K
* * * * * *2
t
e,j ,j 1 t 2 / k* (10.7.18)
¶t e

where

c
F = U 2 / L2 F * h . (10.7.19)

Generalizations to include body forces, species equations and so forth are reasonably
straightforward.

© Fluent Inc., Dec-98 10-37


Free-Surface Flows NONDIMENSIONALIZATION

10.8 Free-Surface Flows


In the presence of a free surface, a fluid is in general subject to four distinct forces: vis-
cous, inertial, gravitational and surface tension. Forming ratios of pairs of these forces
leads to the definition of three independent dimensionless parameters:
inertia force rUL
Reynolds number Re = = (10.8.1)
viscous force m

inertia force U2
Froude number Fr = = (10.8.2)
gravitational force gL
viscous force mU
Capillary number Ca = = (10.8.3)
surface tension force g

where g is the surface tension. Other parameters can be formed that are sometimes
useful, but in the absence of other forces, are merely combinations of the above. The most
common of these is probably the Bond number, Bo, defined by
2
rgL
Bo = (10.8.4)
g

which is a ratio of gravitational force to surface tension force. Another is the Weber
number, We, defined by
2
rLU
We = (10.8.5)
2g

which is a ratio of inertia force to surface tension force.


Of the three parameters Re, Fr, and Ca, the first two were introduced and utilized in
Section 10.2, above. The capillary number, Ca, is the new additional parameter that re-
lates specifically to free-surface flows, in particular surface tension effects.
For a purely isothermal flow the equation of momentum conservation is

FG ¶u + u u IJ = s
H ¶t K + rgi
i
r j i, j ij , j . (10.8.6)

Suppose that G is a segment of the boundary of the computational domain which is a free
surface. On this boundary three boundary conditions have to be applied: the kinematic
constraint, the tangential stress boundary condition and the normal stress boundary condi-
tion. However, in the case that the surface tension is constant the first two of these are
invariant to nondimensionalization, and so only the third, the normal stress condition, is
of interest here.

10-38 © Fluent Inc., Dec-98


NONDIMENSIONALIZATION Free-Surface Flows

This condition has the form


s ij n j b
= - pa + 2g H ni g on G (10.8.7)

where pa is ambient pressure and H is the mean curvature of the surface. The nondimen-
sionalizations available to us are those we have been discussing under the headings of
format (a) and format (b) throughout this chapter. In case we choose format (a) we set
s ij
*
= s ij rU
2
, pa* = pa rU 2 , H * = HL (10.8.8a)

and other quantities are scaled as before. This leads to the dimensionless momentum
equation
*
¶ui
*
+ u*j ui*, j = s *ij , j + Fr -1 gi* (10.8.9a)
¶t

with boundary condition


1
s ij
*
n j = - pa* + 2H * . (10.8.10a)
Re Ca
If you select format (b) you will set
s ij
**
a
= L mU fs ij a f
, pa** = L mU pa , H * = HL (10.8.8b)

which leads to the momentum equation

FG ¶u*
IJ
H ¶t K
+ u*j ui*, j = s *ij , j + Re Fr -1 gi*
i
Re *
(10.8.9b)

with boundary condition


1
s
**
ij n j = - pa** + 2H * . (10.8.10b)
Ca
As was the case in Section 10.2, we recommend that format (a) be used for high Reynolds
number flows and format (b) for low Reynolds number flows. Note that the value of the
surface tension to be entered on the SURFACETENSION command is 1/ReCa if format
(a) is used and 1/Ca if format (b) is used. Most practical free-surface problems are low
Reynolds number, so it will be more likely that you use format (b). The solution method
you choose to adopt for such problems is very significantly dependent on whether the
capillary number is large or small (see Chapter 6 of this manual).
When temperature or concentration gradients are present, important phenomena are
known to occur as a result of the variation of surface tension with temperature or species

© Fluent Inc., Dec-98 10-39


Free-Surface Flows NONDIMENSIONALIZATION

concentration. To illustrate how nondimensionalization should be performed in this case,


consider a situation where the surface tension varies with temperature
g =g Taf . (10.8.11)

As we did with variable density problems we need to select a reference value g 0 of the
surface tension, and also a coefficient g T of surface tension variation which is a measure
of the rate of change of the surface tension with temperature:

FG Dg IJ
g T =O
H DT K . (10.8.12)

The introduction of the new physical quantity g T leads to the definition of a new dimen-
sionless parameter, the Marangoni number Ma, defined by
g T DT
Ma = . (10.8.13)
mU

The boundary conditions at the free surface are conditions on the normal and tangential
components of stress, and are
s ij n j ni b
= - pa + 2g H g (10.8.14)

¶g
s ij n j ti = g ,i ti = T,i ti , (10.8.15)
¶T

respectively. The nondimensionalizations of g and its derivative should be expressed as

g = g 0g * ,
¶g
¶T
= g TF T* c h (10.8.16)

where, of course, F = 1 if g means a linear function of T.

Using the same scalings as before you will obtain either (10.8.10a) or (10.8.10b) for the
normal stress equation (10.8.14). For the tangential stress equation you will obtain,

*
s ij n j ti =
Ma
Re
c h
F T * T,i*ti (10.8.17a)

or
*
s ij n j ti = MaF T * T,i*ti c h . (10.8.17b)

The momentum equations remain unchanged, and the temperature equations are as out-
lined in Section 10.3.

10-40 © Fluent Inc., Dec-98

You might also like