Unit 5 With QP-1
Unit 5 With QP-1
Turbulence models, mixing length model, Two equation (k-Є) models – High and
low Reynolds number models – Structured Grid generation – Unstructured Grid
generation – Mesh refinement –Adaptive mesh – Software tools.
TURBULENCE MODELS
Reynolds decomposition
The random nature of a turbulent flow precludes an economical description of the motion of all
the fluid particles. Instead the velocity in above graph is decomposed into a steady mean value U
with a fluctuating component u′(t) superimposed on it: u(t) = U + u′(t). This is called the Reynolds
decomposition
. A turbulent flow can now be characterized in terms of the mean values of flow properties
(U, V, W, P etc.) and some statistical properties of their fluctuations (u′, v′, w′, p′ etc.)
Even in flows where the mean velocities and pressures vary in only one or two space
dimensions, turbulent fluctuations always have a three dimensional spatial character.
Effective mixing
Particles of fluid which are initially separated by a long distance can be brought close together by
the eddying motions in turbulent flows. As a consequence, heat, mass and momentum are very
effectively exchanged.
For example, a streak of dye which is introduced at a point in a turbulent flow will rapidly
break up and be dispersed right across the flow. Such effective mixing gives rise to high values of
diffusion coefficients for mass, momentum and heat.
Vortex stretching
The largest turbulent eddies interact with and extract energy from the mean flow by process
called vortex stretching
The presence of mean velocity gradients in sheared flows distorts the rotational turbulent
eddies. Suitably aligned eddies are stretched because one end is forced to move faster than the
other.
Energy Cascade
Smaller eddies are themselves stretched strongly by somewhat larger eddies and more weakly
with the mean flow. In this way the kinetic energy is handed down from large eddies to
progressively smaller and smaller eddies in what is termed the energy cascade.
Kolmogorov Microscales
The smallest scales of motion in a turbulent flow (lengths of 0.1 to 0.01 mm and frequencies
around 10 kHz in typical turbulent engineering flows) are dominated by viscous effects.
The Reynolds number Reη of the smallest eddies based on their characteristic velocity υ and
characteristic length η is equal to 1, Reη = υη/ν = 1, so the smallest scales present in a turbulent
flow are those for which the inertia and viscous effects are of equal strength. These scales are
named the Kolmogorov microscales
At these scales work is performed against the action of viscous stresses, so that the energy
associated with small-scale eddy motions is dissipated and converted into thermal internal
energy. This dissipation results in increased energy losses associated with turbulent flows.
Dimensional analysis can be used to obtain ratios of the length, time and velocity scales of
the small and large eddies.
The Kolmogorov microscales can be expressed in terms of the rate of energy dissipation of a
turbulent flow and the fluid viscosity, which uses the notion that in every turbulent flow the
rate of production of turbulent energy has to be broadly in balance with its rate of dissipation
to prevent unlimited growth of turbulence energy.
Flows which possess one or more points of inflexion amplify long wavelength disturbances at
all Reynolds numbers typically above about 10.
After the flow emerges from the orifice the laminar exit flow produces the rolling up of a vortex
fairly close to the orifice. Subsequent amplification involves the formation of a single vortex
of greater strength through the pairing of vortices.
A short distance further downstream, three-dimensional disturbances cause the vortices to
become heavily distorted and less distinct. The flow breaks down, generating a large number
of small-scale eddies, and the flow undergoes rapid transition to the fully turbulent regime.
Mixing layers and wakes behind bluff bodies exhibit a similar sequence of events, leading to
transition and turbulent flow.
A mixing layer forms at the interface of two regions: one with fast and the other with slow
moving fluid. In a jet a region of high-speed flow is completely surrounded by stationary fluid.
A wake is formed behind an object in a flow, so here a slow moving region is surrounded by
fast moving fluid. Figure is a sketch of the development of the mean velocity distribution in
the stream wise direction for these free turbulent flows.
It is clear that velocity changes across an initially thin layer are important in all three flows.
Transition to turbulence occurs after a very short distance in the flow direction from the point
where the different streams initially meet; the turbulence causes vigorous mixing of adjacent
fluid layers and rapid widening of the region across which the velocity changes take place.
It is also immediately clear that the turbulent part of the flow contains a wide range of length
scales. Large eddies with a size comparable to the width across the flow are occurring alongside
eddies of very small size.
The visualization correctly suggests that the flow inside the jet region is fully turbulent, but the
flow in the outer region far away from the jet is smooth and largely unaffected by the
turbulence. The position of the edge of the turbulent zone is determined by the (time-
dependent) passage of individual large eddies. Close to the edge these will occasionally
penetrate into the surrounding region.
In these formulae Umax and Umin represent the maximum and minimum mean velocity at a
distance x downstream of the source. Hence, if these local mean velocity scales are chosen and
x is large enough, the functions f, g and h are independent of distance x in the flow direction.
Such flows are called self-preserving.
Formula (3.16) is called the law of the wall and contains the definitions of two important
dimensionless groups, u+ and y+.
Note: that the appropriate velocity scale is uτ = τw/ρ, the so-called friction velocity.
Velocity-defect law.
Far away from the wall we expect the velocity at a point to be influenced by the retarding
effect of the wall through the value of the wall shear stress, but not by the viscosity itself. The
length scale appropriate to this region is the boundary layer thickness δ.
U = g(y, δ, ρ, τw)
In this region we have the most useful form emerges if we view the wall shear stress as the cause
of a velocity deficit Umax − U which decreases the closer we get to the edge of the boundary layer
or the pipe centre line. Thus
This formula is called the velocity-defect law.
Linear or viscous sub-layer ---(the fluid layer in contact with a smooth wall)
At the solid surface the fluid is stationary and turbulent eddying motions must also stop
very close to the wall and the behavior of the fluid closest to the wall is dominated by viscous
effects. This viscous sub-layer is in practice extremely thin ( y+ < 5) and we may assume that
the shear stress is approximately constant and equal to the wall shear stress τw throughout the
layer. Thus
After some simple algebra and making use of the definitions of u+ and y+ this leads to
Because of the linear relationship between velocity and distance from the wall the fluid layer
adjacent to the wall is also known as the linear sub-layer.
Outer layer --- the inertia-dominated region far from the wall
Experimental measurements show that the log-law is valid in the region 0.02 < y/δ < 0.2. For
larger values of y the velocity-defect law provides the correct form.
In the overlap region the log-law and velocity defect law have to be equal.
where A is a constant. This velocity-defect law is often called the law of the wake.
The turbulent boundary layer adjacent to a solid surface is composed of two regions:
The inner region: 10–20% of the total thickness of the wall layer; the shear stress is (almost)
constant and equal to the wall shear stress τw. Within this region there are three zones. In
order of increasing distance from the wall we have:
– The linear sub-layer: viscous stresses dominate the flow adjacent to surface
– The buffer layer: viscous and turbulent stresses are of similar magnitude
– The log-law layer: turbulent (Reynolds) stresses dominate.
The outer region or law-of-the-wake layer: inertia-dominated core flow far from wall; free
from direct viscous effects.
Turbulence models
A turbulence model is a computational procedure to close the system of mean flow equations.
For most engineering applications it is unnecessary to resolve the details of the turbulent
fluctuations.
Turbulence models allow the calculation of the mean flow without first calculating the full
time-dependent flow field.
Turbulence causes the appearance in the flow of eddies with a wide range of length and time
scales that interact in a dynamically complex way. Given the importance of the avoidance or
promotion of turbulence in engineering applications, it is no surprise that a substantial amount
of research effort is dedicated to the development of numerical methods to capture the
important effects due to turbulence. The methods can be grouped into the following three
categories:
Large eddy simulation: this is an intermediate form of turbulence calculations which tracks
the behaviour of the larger eddies. The method involves space filtering of the unsteady Navier–
Stokes equations prior to the computations, which passes the larger eddies and rejects the
smaller eddies. The effects on the resolved flow (mean flow plus large eddies) due to the
smallest, unresolved eddies are included by means of a so-called sub-grid scale model.
Unsteady flow equations must be solved, so the demands on computing resources in terms of
storage and volume of calculations are large, but (at the time of writing) this technique is
starting to address CFD problems with complex geometry.
Direct numerical simulation (DNS): these simulations compute the mean flow and all
turbulent velocity fluctuations. The unsteady Navier–Stokes equations are solved on spatial
grids that are sufficiently fine that they can resolve the Kolmogorov length scales at which
energy dissipation takes place and with time steps sufficiently small to resolve the period of
the fastest fluctuations. These calculations are highly costly in terms of computing resources,
so the method is not used for industrial flow computations.
Boussinesq hypothesis
Many turbulence models are based upon the Boussinesq hypothesis.
It was experimentally observed that turbulence decays unless there is shear in
isothermal incompressible flows.
Turbulence was found to increase as the mean rate of deformation increases.
Boussinesq proposed in 1877 that the Reynolds stresses could be linked to the
mean rate of deformation.
Using the suffix notation where i, j, and k denote the x-, y-, and z-directions respectively,
viscous stresses are given by:
ui u j
ij eij
x j xi
U i U j
ij ui ' u j ' t
x j xi
Turbulent viscosity
A new quantity appears: the turbulent viscosity t.
U i U j
ij ui ' u j ' t
x j xi
It is, however, assumed to be isotropic. It is the same in all directions. This assumption is
valid for many flows, but not for all (e.g. flows with strong separation or swirl).
RSNS equation is focused on the mean flow and the effects of turbulence on mean flow
properties. Prior to the application of numerical methods the Navier–Stokes equations are time
averaged
Extra terms appear in the time-averaged (or Reynolds averaged) flow equations due to the
interactions between various turbulent fluctuations. These extra terms are modelled with
classical turbulence models: among the best known ones are the k–ε model and the Reynolds
stress model. The computing resources required for reasonably accurate flow computations are
modest, so this approach has been the mainstay of engineering flow calculations over the last
three decades
The vast majority of turbulent flow computations has been and for the foreseeable future will
continue to be carried out with procedures based on the Reynolds-averaged Navier–Stokes
(RANS) equations.
In order to be able to compute turbulent flows with the RANS equations it is necessary to
develop turbulence models to predict the Reynolds stresses and the scalar transport terms and
close the system of mean flow equations.
The time-averaging operation is itself an integration. Thus, the order of time averaging and
summation, further integration and/or differentiation can be swapped or commuted, so this is called
the commutative property.
Since div and grad are both differentiations, the above rules can be extended to a fluctuating vector
quantity a = A + a′ and its combinations with a fluctuating scalar ϕ = Φ +φ′:
To start with we consider the instantaneous continuity and Navier–Stokes equations in a Cartesian
co-ordinate system so that the velocity vector u has x-component u, y-component v and z-
component w:
The effects of fluctuations on the mean flow using the Reynolds decomposition in equations (3.23)
and (3.24a–c) and replace the flow variables u (hence also u, v and w) and p by the sum of a mean
and fluctuating component. Thus
It is important to note that the terms (I), (II), (IV) and (V) in (3.26a–c) also appear in the
instantaneous equations (3.24a–c), but the process of time averaging has introduced new terms
(III) in the resulting time-average momentum equations.
The equation set (3.25) and (3.27a–c) is called the Reynolds-averaged Navier–Stokes equations.
On dimensional grounds we assume the kinematic turbulent viscosity ν , which has dimensions m2/s, can tbe
expressed as a product of a turbulent velocity scale ϑ (m/s) and a turbulent length scale (m). If one velocity scale
and one length scale suffice to describe the effects of turbulence, dimensional analysis yields
ν = Cϑ (3.36)
t
where C is a dimensionless constant of proportionality. Of course the dynamic turbulent viscosity is given
by
µ = Cρϑ
t
Most of the kinetic energy of turbulence is contained in the largest eddies, and turbulence length scale is therefore
characteristic of these eddies which interact with the mean flow. If we accept that there is a strong
connectionbetween the mean flow and the behaviour of the largest eddies we can attempt to link the characteristic
velocity scale of the eddies with the mean flow properties. This has been found to work well in simple two-
dimensional turbulent flows where the only significant Reynolds stress is τxy = τyx = −ρu′v′and the only
significant mean velocity gradient is ∂U/∂y. For such flows it is at least dimensionally correct to state that, if the
eddy length scale is ,
∂U
ϑ=c (3.37)
∂y
where c is a dimensionless constant. The absolute value is taken to ensure that the velocity scale is always a
positive quantity irrespective of the sign of the velocity gradient.
Combining (3.36) and (3.37) and absorbing the two constants C and c into a new length scale m we obtain
This is Prandtl’s mixing length model. Using formula (3.33) and noting that ∂U/∂y is the only significant
mean velocity gradient, the turbulent Reynolds stress is described by
∂U ∂U
τxy = τyx = −ρ u′v′ = ρm2 (3.39)
∂y ∂y
The mixing length model can also be used to predict turbulent transport of scalar quantities. The only
turbulent transport term which matters in the two-dimensional flows for which the mixing length is useful is
modelled as follows:
∂Φ
−ρ v′ϕ′ = Γt (3.40)
∂y
where Γt = µ /σ and µ = ρνt where ν is found from (3.38). Rodi (1980) recommended values for σ of 0.9 in near-
wall flows, 0.5 for jets and mixing layers and 0.7 in axisymmetric jets.
The mixing length has been found to be very useful in simple two-dimensional flows with slow changes in
the flow direction. In these cases the production of turbulence is in balance with its dissipation throughout the
flow, and turbulence properties develop in proportion with a mean flow length scale L. This means that in such
flows the mixing length m is pro-portional to L and can be described as a function of position by means of a
simple algebraic formula. The majority of practically important flows, however, involve additional contributions
to the budgets of turbulence properties due to transport, i.e. convection and diffusion. Moreover, the
production and destruction processes may be modified by the flow itself. Consequently, the mixing length
model is not used on its own in general-purpose CFD, but we will find it embedded in many of the more
sophisti-cated turbulence models to describe near-wall flow behaviour as part of the treatment of wall boundary
conditions.
Advantages:
• easy to implement and cheap in terms of computing resources
• good predictions for thin shear layers: jets, mixing layers, wakes and boundary layers
• well established
Disadvantages:
• completely incapable of describing flows with separation and recirculation
• only calculates mean flow properties and turbulent shear stress
The k---ε model
In two-dimensional thin shear layers the changes in the flow direction are always so slow that the turbulence
can adjust itself to local conditions. In flows where convection and diffusion cause significant differences
between produc-tion and destruction of turbulence, e.g. in recirculating flows, a compact algebraic
prescription for the mixing length is no longer feasible. The way forward is to consider statements regarding
the dynamics of turbulence. The k–ε model focuses on the mechanisms that affect the turbulent kinetic
energy. Some preliminary definitions are required first. The instantaneous kinetic energy k(t) of a turbulent
flow is the sum of the mean kinetic energy Κ =
–12 (U2 + V 2 + W 2) and the turbulent kinetic energy k = –12 ( u′2 + v′2 + w′2 ):
k(t) = K + k
In the developments below we extensively need to use the rate of deforma-tion and the turbulent stresses.
To facilitate the subsequent calculations it is
common to write the components of the rate of deformation sij and the stresses τij in tensor (matrix) form:
∂U ∂u′ ∂V ∂ v′
sxx(t) = Sxx + s′xx = + syy(t) = Syy + s′yy = +
∂x ∂x ∂y ∂y
∂W ∂w′
szz(t) = Szz + s′zz = +
∂z ∂z
1 G ∂U ∂V J 1 G ∂ u′ ∂ v′ J
sxy(t) = Sxy + s′xy = syx(t) = Syx + s′yx = H + K + H + K
2 I ∂y ∂x L 2 I ∂ y ∂x L
1 G ∂U ∂W J 1 G ∂ u′ ∂ w′ J
sxz(t) = Sxz + s′xz = szx(t) = Szx + s′zx = H + K + H + K
2 I ∂z ∂x L 2 I ∂z ∂x L
1 G ∂V ∂W J 1 G ∂ v′ ∂ w′ J
syz(t) = Syz + s′yz = szy(t) = Szy + s′zy = H + K + H + K
2 I ∂z ∂y L 2 I ∂z ∂y L
The product of a vector a and a tensor bij is a vector c whose components can be calculated by application of
the ordinary rules of matrix algebra:
Gb11 b12 b13J Ga1b11 + a2b21 + a3b31JT Gc1JT
abij ≡ aibij = [a1 a2 a3] Hb21 b22 b23K = Ha1b12 + a2b22 + a3b32K = Hc2K = cj = c
Ib31 b32 b33L Ia1b13 + a2b23 + a3b33L Ic3L
The scalar product of two tensors aij and bij is evaluated as follows:
aij . bij = a11b11 + a12b12 + a13b13 + a21b21 + a22b22 + a23b23
+ a31b31 + a32b32 + a33b33
We have used the convention of the suffix notation where the x-direction is denoted by subscript 1, the y-
direction by 2 and the z-direction by 3. It can be seen that products are formed by taking the sum over all
possible values of every repeated suffix.
Or in words
The transport terms (III), (IV) and (V) are all characterised by the appearance of div and it is common
practice to place them together inside one pair of brackets. The effects of the viscous stresses on K have
been split into two parts: term (IV), the transport of K due to viscous stresses; and term (VI), the viscous
dissipation of mean kinetic energy K. The two terms that contain the Reynolds stresses −ρ ui′uj′
account for turbulence effects: term (V) is the turbulent transport of K by means of Reynolds stresses and
(VII) is the net decrease of K due to deformation work by Reynolds stresses giving rise to turbulence
production. In high Reynolds number flows the turbulent terms (V) and (VII)VII) are always much larger
than their viscous counterparts (IV) and
∂ (ρk)
+ div(ρkU) = div(− p′u′ + 2µ u′s′ij − ρ –12 ui′ . ui′uj′) − 2µ s′ij . s′ij − ρui′uj′ . Sij (3.42)
∂t
(I) (II) (III) (IV) (V) (VI) (VII)
In words
Equations (3.41) and (3.42) look very similar in many respects; however, the appearance of primed quantities
on the right hand side of the k-equation shows that changes to the turbulent kinetic energy are mainly governed
by turbulent interactions. Terms (VII) in both equations are equal in magni-tude, but opposite in sign. In two-
dimensional thin shear layers we found (see section 3.4) that the only significant Reynolds stress −ρ u′v′ is
usually positive if the main term of Sij in such a flow, the mean velocity gradient ∂U/∂y, is positive. Hence term
(VII) gives a positive contribution in the k-equation and represents a production term. In the K-equation,
however, the sign is negative, so there the term destroys mean flow kinetic energy. This expresses
mathematically the conversion of mean kinetic energy into turbulent kinetic energy.
The k---εε model equations
It is possible to develop similar transport equations for all other turbulence quantities including the rate of
viscous dissipation ε (see Bradshaw et al., 1981). The exact ε-equation, however, contains many unknown and
unmea-surable terms. The standard k–ε model (Launder and Spalding, 1974) has two model equations, one
for k and one for ε, based on our best understand-ing of the relevant processes causing changes to these
variables.
We use k and ε to define velocity scale ϑ and length scale representative of the large-scale turbulence as
follows:
k3/2
ϑ = k1/2 =
ε
One might question the validity of using the ‘small eddy’ variable ε to define the ‘large eddy’ scale . We are
permitted to do this because at high Reynolds numbers the rate at which large eddies extract energy from the
mean flow is broadly matched to the rate of transfer of energy across the energy spectrum to small, dissipating,
eddies if the flow does not change too rapidly. If this was not the case the energy at some scales of turbulence
could grow or diminish without limit. This does not occur in practice and justifies the use of ε in the definition
of .
Applying dimensional analysis we can specify the eddy viscosity as follows:
k2
µt = Cρϑ = ρCµ (3.44)
ε
∂ (ρk) Gµ J
+ div(ρkU) = div H t grad kK + 2µtSij . Sij − ρε (3.45)
∂t I σk L
∂ (ρε) Gµ J ε ε2
+ div(ρεU) = div H t grad εK + C1ε 2µtSij . Sij − C2ε ρ (3.46)
∂t I σε L k k
The equations contain five adjustable constants: Cµ, σk, σε, C1ε and C2ε. The standard k–ε model employs
values for the constants that are arrived at by comprehensive data fitting for a wide range of turbulent flows:
A ∂U ∂Uj D 2 2
−ρ ui′uj′ = µt B i + E − ρkδij = 2µtSij − ρkδij (3.48)
∂
C jx ∂xiF 3 3
Boundary conditions
The model equations for k and ε are elliptic by virtue of the gradient diffu-sion term. Their behaviour is similar to the
other elliptic flow equations, which gives rise to the need for the following boundary conditions:
At high Reynolds number the standard k–ε model (Launder and Spalding, 1974) avoids the need to integrate
the model equations right through to the wall by making use of the universal behaviour of near-wall flows discussed
in section 3.4. If y is the co-ordinate direction normal to a
solid wall, the mean velocity at a point at yP with 30 < yP+ < 500 satisfies the log-law (3.19), and measurements of
turbulent kinetic energy budgets indi-cate that the rate of turbulence production equals the rate of dissipation. Using
these assumptions and the eddy viscosity formula (3.44) it is possible to develop the following wall functions, which
relate the local wall shear stress (through uτ) to the mean velocity, turbulence kinetic energy and rate of dissipation:
U 1 uτ2 uτ3
u+ = = ln(EyP+) k= ε= (3.49)
uτ κ Cµ κy
Von Karman’s constant κ = 0.41 and wall roughness parameter E = 9.8 for smooth walls. Schlichting (1979) also
gives values of E that are valid for rough walls.
For heat transfer we can use a wall function based on the universal near-wall temperature distribution valid at
high Reynolds numbers (Launder and Spalding, 1974)
(T − Tw)Cpρuτ A Gσ JD
T+ ≡ − = σT,t B u+ + P H T,l K E (3.50)
qw C I σT,t L F
Finally P is the pee-function, a correction function dependent on the ratio of laminar to turbulent Prandtl numbers
(Launder and Spalding, 1974).
At low Reynolds numbers the log-law is not valid, so the above-mentioned boundary conditions cannot be used.
Modifications to thek–ε model to enable it to cope with low Reynolds number flows are reviewed in Patel et al.
(1985). Wall damping needs to be applied to ensure that viscous
and
η(1 − η/η0) k
C 1*ε = C1ε − η= 2Sij . Sij η0 = 4.377 β = 0.012
1 + βη3 ε
Only the constant β is adjustable; the above value is calculated from near-wall turbulence data. All other constants are
explicitly computed as part of the RNG process.
The ε-equation has long been suspected as one of the main sources of accuracy limitations for the standard version
of the k–ε model and the RSM in flows that experience large rates of deformation. It is, therefore, interest-ing to note
that the model contains a strain-dependent correction term in the
constant C1ε of the production term in the RNG model ε-equation (it can also be presented as a correction to the sink
term).
11.5 Block-structured
grids To overcome the problems associated with structured grid generation for
complex geometries, block-structured CFD methods have been developed. In a
block-structured grid, the domain is sub-divided into regions, each of which
has a structured mesh. The mesh structure in each region can be different, and
it is even possible to use different co-ordinate systems. Such meshes are more
flexible than (‘single block’) structured meshes described in the previous
sections. The block-structured approach allows the use of fine grids in regions
where greater resolution is required. The interfaces of adjacent blocks may have
grids on either side that are matching or non-matching, but, either way, they
must be properly treated in a fully conserva-tive manner. In some codes the
solvers are applied in a block-wise manner (block by block with overall final
iterations to unify boundary conditions) and local refinement is possible block-
wise. Block-structured grids with overlapping regions are called composite grids
or chimera grids. Figure 11.9 shows a Cartesian block-structured grid used for
the calculation of flow over an aerofoil. The resulting grid structure combines
the advantages of Cartesian grids – easy to generate, equations simple to
discretise and solve –with the ability of curvilinear grids to accommodate curved
complex bound-aries (see Courier and Powell, 1996).
In unstructured grid arrangements we are not restricted to one particular cell type, but it is possible to use a mixture of
cell shapes. In 2D a mixture of triangular and quadrilateral elements can be used to construct the grid. In 3D flow
calculations combinations of tetrahedral and hexahedral elements are frequently used. Such grids are called hybrid
meshes. Figure 11.12 shows an example of a hybrid unstructured grid for the calculation of flow in a tube bank where
quadrilateral cells have been used near solid walls to provide better resolution of the viscous effects in the boundary layers
and an expanding triangular mesh structure elsewhere to utilise the resources efficiently.
The most attractive feature of the unstructured mesh is that it allows the calculation of flows in or around geometrical
features of arbitrary complexity without having to spend a long time on mesh generation and mapping. Grid generation
is fairly straightforward (especially with triangu-lar and tetrahedral grids), and automatic generation techniques,
originally developed for finite element methods, are now widely available. Further-more, mesh refinement and adaption
(semi-automatic mesh refinement to improve resolution in regions with large gradients) are much easier in unstructured
meshes. In the sections to follow we explore the unstructured methodology in more detail as it is now the most popular
technique and is included in all commercial CFD codes on the market today.
11.7 Discretisation in
unstructured Unstructured grids are the most general form of grid arrangement for
grids most complex geometries. Here we present a brief outline of discretisation
techniques for unstructured grids with arbitrary cell shapes, which may be
bounded by any number of control surfaces. We limit ourselves to the devel-
opment of the main ideas; interested readers should consult the literature for
further details of the methodology.
There are two ways of defining control volumes in unstructured meshes:(i) i)
cell-centred control volumes and (ii) vertex-centred control volum These two
variants are illustrated in Figure 11.13 for a 2D problem.
In the cell-centred method the nodes are placed at the centroid of the
control volume, as shown in Figure 11.13a. In the vertex-centred method the
nodes are placed on the vertices of the grid. This is followed by a process known
as median-dual tessellation, whereby sub-volumes are formed by joining
centroids of the elements and midpoints of the edges, as illustrated in
Figure 11.13b. The sub-volume surrounding a node then forms the control volume for discretisation. Both cell-centred and
vertex-centred methods are used in practice. We develop the ideas of discretisation in unstructured grids for the cell-centred
method, which is simpler to understand, and, since a control volume always has more vertices than centroids, it has slightly
lower storage requirements than the vertex-centred method.
Principles of Grid
Generation
{ = i/imax, i = 0, 1, 2, ·· ·, im ax
TJ=i/imax, j=0,1,2,··•,jmax (11.1)
(=k/kmax, k=0, 1, 2, · · ·, km ax
C-Grid Topology
In the case of the C-topology the aerodynamic body is enclosed by one family of
grid lines, which also form the wake region (if present). The situation is sketched
in Fig. 11.5. As we can see, the lines T/ = const. start at the farfield ({ = 0),
follow the wake, pass the trailing edge (node b), wrap in clockwise direction
round the body, and finally continue to the farfield again ({ = 1). The other
family of grid lines ({ = const.) emanates in normal direction from the body and
the wake. The part (segment) a-b of the grid line T/ = 0 represents a coordinate
cut. This means that the segment a-b in the physical space is mapped onto
two segments in the computational space, namely a � { � b and b' � { � a'.
Hence, the nodes on the upper (b1 -a1) and the lower part (a-b) of the cut are
hold separately in the computer memory. The appropriate boundary condition
was presented in Section 8.6.
Chapter 11
TJ= l ri
f
d e
-
f
- --- -- ,-
- --- - - •f -
__ ,_
e
- -- - - -
ri=O
- - - -- .
0-Grid Topology
We can see from the rendering of the O-topology in Fig. 11.9 that one family of
grid lines (ri = const.) forms dosed curves around the aerodynamic body. The
second family of grid lines (( = canst.) is spanned in radial direction between
the body and the outer boundary (farfield in this case). The complete boundary
line T/ = 0 represents the contour of the body (from a to a'). The coordinate
cut is defined by the boundaries ( = 0 (nodes a-c) and ( = 1 (nodes a'-c') in
the computational space. The example in Fig. 11.10 shows a standard O-grid
used to simulate inviscid flow past an airfoil. A disadvantage of the O-topology
is the poor grid quality at a sharp trailing edge.
1
Adaptive grid
An adaptive grid is a grid network that automatically clusters grid points in
regions of high flow-field gradients; it uses the solution of the flow-field properties
to locate the grid points in the physical plane. An adaptive grid can be visualized as
one which evolves in steps of time in conjunction with a time-dependent solution of
the governing flow-field equations, which computes the flow-field variables in steps
of time. During the course of the solution, the grid points in the physical plane move in
such a fashion to "adapt" to regions of large flow-field gradients as these
gradients evolve with time. Hence, the actual grid points in the physical plane are
constantly in motion during the solution of the flow field and become stationary
only when the flow solution approaches a steady state. Therefore, unlike the
stretched grid discussed in Sec. 5.6 and the elliptic grid generation discussed in Sec.
5.7, where the generation of the grid is completely separate from the flow-field
solution, an adaptive grid is intimately linked to the flow-field solution and changes
as the flow field changes. The hoped-for advantages of an adaptive grid are
associated with the grid points being automatically clustered in regions where the
"action" is occurring. These advantages are (1) increased accuracy for a fixed
number of grid points or (2) for a given accuracy, fewer grid points are needed.
Adaptive grids are still very new in CFD, and whether or not these advantages are
always achieved is not well established.
An example of a simple adaptive grid is that used by Corda (Ref. 35) for the
solution of viscous supersonic flow over a rearward-facing step. Here, the trans
formation is expressed in the form
2 and 16 Marks