This document is meant to introduce the basic routes employed within TransAT for the modelling and
simulation of turbulent fluid flow and scalar convection, without exploring the foundations and model
derivations. We briefly present the transport equations and models for RANS and Scale-resolving
strategies, including LES and V-LES. The document presents in addition a thorough description of near
wall treatment on both approaches. Only the models that have been validated are presented in this
document. The extension of these models for multiphase turbulent flows is not treated here. Examples of
applications of the models and strategies can be found in internal reports available on the
webpage.
In this Chapter, the Reynolds-averaged equations for all quantities used in TransAT are presented. The
initial equations, on which Reynolds averaging is applied, are presented in the Equations and Algorithmsmanual, Chapter Incompressible Navier-Stokes Equations .
1.1 Reynolds Averaging Procedure
According to the Reynolds flow-decomposition procedure, a flow variable ϕ can be expressed as the sum
of an average and a fluctuating quantity:
where ϕ is further taken to be the ensemble average of ϕ and ϕ′ is the fluctuation around this mean
(ϕ′ = 0). Before proceeding further, it is necessary to recall the statistical rules of Reynolds averaging,
often referred to as Reynolds conditions, which any two flow variables ϕ and ψ must obey:
and
1.2 The Reynolds Averaged Navier-Stokes Equations
For constant density and viscosity, Reynolds averaging, applied to the system of instantaneous equations
of conservation yields the Reynolds averaged mass and momentum conservation equations, known as the
Reynolds Averaged Navier-Stokes Equations (RANS):
where τij≡u′iu′j stands for the Reynolds-stress tensor (which is symmetric, i.e. τij = τji ) representing
the contribution of the turbulent motion generated by velocity fluctuations to the mean stress
tensor in momentum equations. This turbulent motion will in turn result in an increase of
momentum exchange and mixing. The off-diagonal components of τij , or shear stresses
(u′v′,v′w′, and u′w′), prevail in theory in the transport of mean momentum by turbulent
motion, while the diagonal terms, or normal stresses (u′2,v′2 , and w′2 ), only play a minor
role.
1.3 The Reynolds Averaged Energy Equation
The equation of energy conservation can be averaged with respect to time and solved together with the
equations describing the mean flow and turbulence. If this equation is written for the temperature, the
Reynolds-averaging procedure applied to T yields
where q′′total = q′′j + Q′′j is the total rate of heat transfer due to both molecular and turbulent
motions, and Q′′j = -ρcpu′jθ′ represents the turbulent heat-flux.
To solve the system of RANS equations 1.4, 1.5 and 1.6, where each barred quantity represents a
Reynolds average, we can take two avenues: Either the turbulent stresses and heat fluxes
are determined individually by solving a set of transport equations for each component (a
total of six + three), or one introduces proper closure relations for τij and the turbulent
heat-fluxQ′′j . The closure relations must provide a physically coherent representation of turbulence
mechanisms.
1.4 The Reynolds Averaged Scalar Equation
In the same way as Navier-Stokes equations, the scalar transport equations can be averaged using the
Reynolds-averaging procedure. This gives the following equation
where is the molecular diffusivity, u′c′ is the turbulent scalar transport and FC is the averaged source
term applied to the scalar.
1.5 The Reynolds Stress Equations
The Reynolds stress, τij can also be determined by solving its own transport equation. A special
notation is introduced for this purpose, namely the Navier-Stokes operator (u′i) producing the
fluctuating–momentum equations
This operator helps to directly derive an equation for the Reynolds stress tensor through construction
of the following second order moment:
After some calculations, one arrives at the Reynolds-stress transport equations:
where
designate, respectively, the mechanical production of turbulence Pij , or the source of Reynolds stress,
the pressure-strain correlation or inter-component transfer term Πij , the dissipation (by the action of
viscosity) rate tensor εij , and the turbulent transport term (Cijk), acting in addition to the molecular
diffusion of τij represented by ν ∇2τij .
The pressure-strain term Πij , gives rise to isotropy of the turbulence field by redistributing the
turbulent kinetic energy among its three components and by reducing the shear stresses (Hinze, 1975).
1.6 The Dissipation Term
A typical practice in modelling the energy dissipation term εij (Eq. 1.13) is to assume that the
dissipative eddies are isotropic (Hinze, 1975) and thus
This hypothesis, known as the Kolmogorov assumption of local isotropy, supposes that small turbulent
structures associated with the rate of dissipation do not have a preferential direction. First, a transport
equation for the dissipation rate tensor εij has to be established. For the sake of brevity we restrict
ourselves to the exact transport equation for ε itself (not for εij ), which can formally be obtained by
constructing the moment:
The equation takes the form
where
correspond to the following physical processes: The production by the mean motion (i.e by the mean
velocity gradients) Pε1 , the generation rate of vorticity fluctuations through the self-stretching of
vortex tubes Pε2 , the decay or destruction of the dissipation Φε through the action of viscosity, and
the turbulent diffusion of dissipation ε (diffusion by molecular effects is represented by
ν∇2ε).
1.7 The Turbulent Kinetic Energy
A modelled version of the exact transport equation (1.10) can be derived, by letting k ≡ τii∕2:
In this equation, ≡ Pii∕2 represents the mechanical production of turbulence due to the interaction
between turbulent stresses and mean velocity gradients. Note also that the contraction of τij yields
Πii = 0 by virtue of u′i,i = 0.
1.8 Chemical species equations
Chemical species transport equations are modelled using their respective mass fraction Yk, where k
represents the species index. When the species have different densities, the usual Reynolds average gives
new unclosed terms ρ′uj′ in the mass balance equation corresponding to the correlation between density
and velocity fluctuations. To avoid this difficulty, mass-weighted averages (called Favre averages) are
usually preferred
(1.23)
where f is a generic quantity. Splitting Yi into a Favre mean and a fluctuation Yi′′ the transport
equation can be written
(1.24)
where Dk is the molecular diffusion coefficient for species k and is the Favre average of the source
term.
Chapter 2 Two-Equation Turbulence Models
There are different ways to model the Reynolds–stress tensor τij (when the transport equations of
τij are not solved explicitly). The best-known approach and the most practical one is based on the eddy
viscosity concept or Boussinesq Hypothesis, which assumes that the Reynolds stress is decomposed into
an isotropic and a deviatoric part,
where δij = 1 for i = j, and zero otherwise. The deviatoric part (2nd term in the RHS of Eq. (2.1)) is a
symmetric, traceless tensor and links τij linearly to the rate–of–strain tensor Sij . The coefficient of
proportionality, νt , designates the eddy viscosity (or turbulent viscosity), which is a characteristic of
the flow.
2.1 The Standard k - ε Turbulence Model
In the k -ε model (Launder & Spalding, 1974), the local state of turbulence is characterised by kinetic
energy k and its dissipation rate ε. The combination of ℓ0≈ k3∕2∕ε and τ0≈ k∕ε (or ℓ0 and vI≡)
leads to the generalised form of the isotropic eddy viscosity νt :
where Cμ is a constant. The distribution of turbulent kinetic energy k and its rate of dissipation ε
appearing in this relation are determined from the following model transport equations:
The reader should refer to Launder & Spalding (1974) for a better understanding of the
derivation of the model from the exact transport equations. The empirical constants are
assigned the following standard values: Cμ = 0.09; Cε1 = 1.44; Cε2 = 1.92; σk = 1. and
σε = 1.3. The standard model (Eqs. 2.2 - 2.4) with its original coefficients is valid only in
flow regions away from solid boundaries (high-Re-number flow regions), and depending on
whether it is to be employed in low- or high-Re forms, special near-wall treatments are
needed..
2.2 The RNG k - ε Turbulence Model
The RNG k - ε model, is more advanced than standard model and was derived from renormalization
group theory. The distribution of turbulent kinetic energy k and its rate of dissipation ε
appearing in this relation are determined from the following model transport equations:
where η = Sk∕ε; η0 = 4.38; β = 0.012 and S denotes strain rate S = . Effective viscosity
instead of turbulent viscosity is used here:
The empirical constants are assigned the following standard values: Cμ = 0.0845; Cε1 = 1.42;
Cε2 = 1.68; σk = 0.71942 and σε = 0.71942. The RNG model (Eqs. 2.5 - 2.7) with its
original coefficients is valid only in flow regions away from solid boundaries (high-Re-number
flow regions), and depending on whether it is to be employed in low- or high-Re forms,
special near-wall treatments are needed. In the energy/temperature equation a new formula
for heat diffusion coefficient is employed which is also valid only for high-Re-number flow:
In the mass transfer equations/concentration equations a new formula for mass diffusivity is employed:
2.3 Extensions to k - ε Turbulence Model
Depending on the flow conditions, different extensions to the basic k - ε model have been
proposed.
2.3.1 Yap correction
The Yap correction (Yap, 1987) consists of an additional source term of the form ρSε to the right hand
side of the epsilon equation. The source term can be written as,
(2.10)
where,
(2.11)
with yn being the normal distance to the nearest wall.
The Yap correction is active in non-equilibrium flows and tends to reduce the departure of the
turbulence length scale from its local equilibrium level. Yap (1987) showed improved results with the
k - ε model in separated flows when using this extra source term.
2.3.2 Swirl correction
The turbulent eddy viscosity is corrected for strongly swirling flows because the standard k -ε models
tend to overestimate turbulent diffusion for swirling flows. The eddy viscosity is modified using a swirl
correction factor, fsw, as follows (Shih et al., 1997),
(2.12)
where,
(2.13)
and
2.3.3 Compressible correction
Initially turbulence models were designed for low-speed, isothermal flows. The compressibility correction
is devised to deal with additional effects seen for higher Mach number flows, specifically, the effects of
compressibility on the dissipation rate of the turbulence kinetic energy. For free shear flows, this is
exhibited as the decrease in growth rate in the mixing layer with increasing Mach number.
Standard turbulence models do not account for this Mach number dependence, and thus a
compressibility correction is used. For compressible flows, two extra terms, known as the dilatation
dissipation, εd, and the pressure-dilatation occur in the turbulence kinetic energy equation. The
pressure-dilatation term is usually neglected because its contributions have been shown to be small
(Wilcox, 1993).
The dilatation dissipation term is included in addition to the solenoidal, or incompressible,
dissipation, ε. Thus, the effect is that the growth rate of turbulence is inhibited when the
correction is active. Sarkar et al. (1991) modeled the ratio of the dilatation dissipation to
the solenoidal dissipation, εd∕ε, as a function of the turbulence Mach number, Mt, defined
as
(2.14)
where c is the speed of sound. For the model by Sarkar et al. (1991), the additional source term to the
k equation is,
(2.15)
Chapter 3 Near–Wall Modelling
3.1 Near–Wall Treatment in High–Re Flows
The near-wall treatment to be presented next applies strictly to the standard k -ε model (Eqs. 2.2
- 2.4), Their combination is referred to as the high-Re k - ε model. This near-wall modelling
consists in bridging the viscous sub layer, i.e. the first grid point must be located outside this
zone.
3.1.1 Standard Wall–Function Approach
At a rigid boundary, the no-slip condition in setting the velocity of the fluid to zero (or to that of the
wall to account for moving-boundaries) is used for laminar flows. In high-Re flows, the wall-function
approach, whose theoretical basis is discussed next, is applied in place of the no-slip condition. This
treatment is based on two assumptions: (i) The immediate near-wall region is in a state of
local equilibrium (= ε), and (ii) the velocity profile merges with the ”log-law” given by
However one needs to make the extra assumption , that, the equilibrium layer near the surface,
where production balances dissipation, is one-dimensional and the stress across is constant. In these
circumstances, it can be demonstrated that
where the subscript p refers to the first control volume centre relative to the wall. Laboratory data for
flow over smooth surfaces indicate that kp∕uτ2≈ 3.3, and hence Cμ≈ 0.09. In a second step and
making use of (3.2) we can relate the induced retarding wall shear-stress w to the velocity vector p
by
where
with
Furthermore, the diffusive flux of k is equal to zero at the wall. In a second step, the wall shear-stress
(supposed to be uniform and equal to that at the wall, y ≥ yw ) acting at a distance yn = 2yp from the
wall induces a rate of turbulence production given by
This quantity (w) has to replace the production term in the transport equation of k (2.3).
Likewise, over the fully turbulent region (yw≤ y ≤ yn ) where k was taken to be uniform (in
the pace interval yw≤ y ≤ dp ), the mean value of εw can be obtained by the expression
It should be noted, however, that since the assumption of ”local equilibrium” near the wall is not
justifiable in the viscous sub layer (yp+< 11.6), the wall-function approach requires the value of yp+
at the cell adjacent to the wall to fall in the range 11.6 ≤ yp+≤ 200 - 500, otherwise, the result is
systematically biased.
3.1.2 Two–layer turbulence models
The two–layer approach adopted here consists of resolving the viscosity–affected regions close to walls
with a one–equation model, while the outer core flow is resolved with the standard k - ϵ model
described above. In the one–equation model, the eddy viscosity is made proportional to a velocity scale
and a length scale lμ. The distribution of lμ is prescribed algebraically while the velocity scale is
determined by solving the k–equation (as in Eq. (2)). The dissipation rate ε appearing as sink term
in the k–equation is related to k and a dissipation length scale lε which is also prescribed
algebraically. The different two-layer versions available in the literature differ in the use
of the velocity scale and the way lμ and lε are prescribed. It should be mentioned that in
the fully turbulent region the length scales lμ and lε vary linearly with distance from the
wall. However, in the viscous sublayer lμ and lε deviate from the linear distribution in order
to account for the damping of the eddy viscosity and the limiting behaviour of ε at the
wall.
The approach combines the standard k - ϵ model in the outer region with a one–equation model
due to Norris and Reynolds Norris & Reynolds (1975) in the viscous–sublayer employing
In this model, the length scale lμ is damped in a similar way as the Prandtl mixing length by the
Van Driest function, so that it involves an exponential reduction governed by the near–wall
Reynolds number Ry = Uyn∕ν. However, in contrast to the original Van Driest function,
Ry uses k1∕2 as a velocity scale U instead of Uτ which can go to zero for separated flows.
The constant Cl is set equal to κCμ-3∕4 to conform with the logarithmic law of the wall. The
empirical constants appearing in the fμ–function are assigned the values Aμ = 50.5 and
A+ = 25. The reader is referred to Rodi (1991) for a review and further details on the choice
of the constants. For the dissipation scale the following distribution is used near the wall:
The outer (k - ϵ) and the near–wall model are matched at a location where the damping function fμ
reaches the value 0.95, i.e. where viscous effects become negligible.
The combination of the Kato–Launder correction with the TLK model is hereafter labeled
TLKK.
The development of this model was motivated by the fact that the length scales-functions as proposed in
Norris & Reynolds (1975) particularly the lε–function, are not in agreement with direct
numerical simulation (DNS) data, and that the normal fluctuations (v′2)1∕2 are a more relevant
velocity scale for the turbulent momentum transfer near the wall than is k1∕2. Therefore, the
following model using (v′2)1∕2 as a velocity scale was proposed in Rodi & Mansour (1993)
which is based on DNS data for fully developed channel flow. As an equation for k is solved, v′2
needs to be related to k, which is done through the following DNS based empirical relation
valid only very near the wall. The matching between the outer and near–wall model is performed at a
location where Ry = 80.
3.2 The Low-Re k - ε Turbulence Model
The standard k - ε model was developed for high Reynolds number flows and is therefore
not valid in flow regions very close to the wall, i.e. within the viscous sub layer. In this
approach, the following relation for νt was proposed; it includes a damping function, fμ ,
varying from almost zero near the wall to ≈ 1 at the outer edge of the viscous sub layer:
Experiments and DNS have also shown that the model constants Cε1 and Cε2 appearing in the source
terms of the transport equation for ε need also to incorporate damping functions f1 and f2 in order
to reproduce the steep gradient of ε near the wall, so that equation (2.4) takes the form
where the molecular diffusion of ε is now re-introduced, unlike in (2.4) for high-Re-number flows. In
this equation Ξ represents an additional term to account for the fact that dissipation processes are not
isotropic within the viscous sub layer (Jones & Launder, 1972). The different low-Re models proposed
so far differ either in the use of the model functions fμ , f1 and f2 and the term Ξ, or the way the
dissipation rate is obtained.
3.2.1 The Jones & Launder and Launder & Sharma Models
The Low-Re k - ε model employed most frequently is due to Jones & Launder (1972). This model is
The model constants Cε1 and Cε2 are assigned the same values as those of the standard k -ε model,
i.e. Cε1 = 1.44 and Cε2 = 1.92 (as in Launder & Sharma (1974)) . This model is found
to result in a very rapid growth of fμ towards 1 (i.e. towards the edge of the viscous sub
layer).
3.2.2 The Abe-Kondoh-Nagano Model
Various new low–Re k - ε have been proposed. One of them is proposed by Abe et al. (1994)
The wall boundary conditions used together with this type of model require the dissipation rate at the
wall ε|wall to be adjusted so as to reproduce the correct asymptotic behaviour, i.e. ε|wall = 2νk∕yp2. In
contrast to the wall–function approach used with the standard k -ε model for high–Re number flows,
in the low-Re schemes the no-slip condition is to be employed for the velocity, along with
the condition of zero turbulent kinetic energy at the wall. Additionally, the model should
meet a criterion similar to the one required for the WF approach, i.e. yp+< 0.1 - 0.5. In
general, a typical layer of about 30 grid–points lying within the viscous sub layer (where
fμ< 0.95) is necessary to correctly reproduce the steep gradient of the dissipation rate.
An alternative approach to the use of the damping function f2 to avoid having to face a singularity
in the destruction term -Cε2ε2∕k consists in replacing this term by
so as to allow the turbulent time scale to recover τ0 = k∕ε far from boundaries – at high–Re
turbulence –, and makes it proportional to the Kolmogorov time scale ~ for low-Re near wall
turbulence (Durbin, 1991, 1993):
where CK is a constant of order one.
3.2.3 The Lam-Bremhorst Model
Lam & Bremhorst (1981) extended the high Reynolds number form of the k - ε model and tested by
application to fully developed pipe flow. The relationship is of the form:
(3.22)
(3.23)
(3.24)
Chapter 4 Turbulent Heat flux Modelling
In order to model the turbulent heat fluxes, the Reynolds Averaged Navier-Stokes equations
(RANS) and the Reynolds Averaged Energy equation introduced in the first chapter need to be
solved.
where τij≡u′iu′j stands for the Reynolds-stress tensor (which is symmetric, i.e. τij = τji ) representing
the contribution of the turbulent motion generated by velocity fluctuations to the mean stress tensor in
momentum equations. This turbulent motion will in turn result in an increase of momentum
exchange and mixing. The off-diagonal components of τij , or shear stresses (u′v′,v′w′,
and u′w′), prevail in theory in the transport of mean momentum by turbulent motion,
while the diagonal terms, or normal stresses (u′2,v′2 , and w′2 ), only play a minor role.
And q′′total = q′′j + Q′′j is the total rate of heat transfer due to both molecular and turbulent motions,
and Q′′j = -ρcpu′jθ′ represents the turbulent heat-flux.
The different modelling approaches for turbulent heat flux modelling are presented in Fig. 4.1
wherein, the simplest approach is called the Simple Gradient Diffusion Hypothesis (SGDH) and the
most complex one is the Turbulent Heat Flux Modelling (THFM) approach which requires the solution
of 5 additional equations to model the turbulent heat flux.
Figure 4.1: Turbulent heat flux modelling overview, where kT represent the variance of the
temperature fluctuations θ′2 and eT represent the dissipation rate of the heat fluxes εθ′.
4.1 Turbulent Heat Flux Model (THFM)
The most accurate approach to account for thermal effects and heat transfer within the time-averaged
context is to solve the transport equation for the turbulent heat fluxes u′iθ′, by incorporating a detailed
modelling of the buoyancy term. In this model, transport equations for the variance of the temperature
fluctuations θ′2 and its dissipation rate εθ′ have to be solved in addition to close the set of
equations.
The modeled transport equations for turbulent heat fluxes are given by: (for i=1,2,3) Carteciano &
Grötzbach (2003)
(4.4)
where κ is the thermal diffusivity: κ = .
The term πi represent the modeling of the pressure-temperature gradient correlation with Monin’s
and Launder’s correlation.
(4.5)
where δi is the Kronecker delta and the index n represent the normal direction to a wall. The dissipation rate of the heat fluxes εu′iθ′ can be modelled by:
(4.6)
Where Ret is the local turbulent Reynolds number, Pet = Ret× Pr is the local Peclet number
and R is the turbulent time-scale ratio. εu′iθ′ is negligible in the transport equation (4.4)
of the heat fluxes at high Peclet numbers but its contribution is important at low Peclet
numbers. Another expression for the dissipation rate is available for very low Peclet numbers:
(4.7)
The buoyancy term Gu′iθ′ is expressed in terms of the variance of the temperature fluctuations
θ′2.
(4.8)
For a detailed description of the buoyancy effects, a transport equation of the variance of the
temperature fluctuations θ′2 is solved:
(4.9)
The previous transport equation requires the dissipation rate of the temperature variance εθ′ to be
modeled using an other transport equation.
with:
(4.11)
4.2 Full Algebraic Turbulent Heat Flux Model: θ′2 - εθ′
In this model, the turbulent heat flux u′iθ′ is modelled algebraically instead of solving an
additional set of partial differential equations. The model is referred to as the Algebraic
Turbulent Heat Flux Model (ATHFM) (Launder, 1988). The turbulent heat flux (u′iθ′) can be
approximated by its production rate times the turbulent time scale. The production terms
and turbulence fluctuations from the differential transport equations are locally in balance
S. Kenjeres (2000).
(4.12)
In order to close the above system, the values of θ′2 and εθ′ are needed. In the case where transport
equations are solved for both the above quantities we obtain the ATHFM -θ′2- εθ′ model.
and
where:
In Kenjereš et al. (2005), the model for the Reynolds stress was extended with a buoyancy and
turbulent heat flux term given as,
with Cθ = 0.15.
This model was further developed and calibrated by Shams et al. (2014) where the eddy viscosity νt
was given as,
Where Red = d∕ν with d being the distance to the wall, and the constants being,
Cμ = 0.09,CT = 0.6,Cd0 = 0.091,Cd1 = 0.0042,Cd2 = 0.00011.
The algebraic turbulent heat flux model used (Shams et al., 2014; Kenjereš et al., 2005) involves
an additional term proportional to the Reynolds stress anisotropy. The algebraic turbulent
heat flux is also redefined with different names for the model constants and is given as,
with gi the gravity vector and aij the Reynolds stress anisotropy tensor defined as,
(4.18)
The transport equation for θ′2 is solved and εθ′ is evaluated from the thermal to mechanical time-scale
ratio (Eq. (4.23)) where an algebraic expression is used by Kenjereš et al. (2005) and a constant value
of 0.5 by Shams et al. (2014). The assumption of the constant time-scale ratio R worked reasonably
well in a number of flow regimes when compared with the respective reference DNS or experimental
data.
In the work of Shams et al. (2014), the model proposed by Kenjereš et al. (2005) is referred to as
AHFM-2005 which was calibrated for natural and mixed convection flow regimes close to unity
Prandtl number. Shams et al. (2014) proposed a new calibration for forced convection and
low Prandtl number fluids (liquid metal) referred to as AHFM-cc. The model was further
extended by introducing a logarithmic dependence of coefficient Ct1 on the Reynolds and
Prandtl numbers with a constraint to be positive for RePr > 180. This variant of AHFM-cc
was called AHFM-NRG. The model constants for the above models are given in Table 4.1.
Ct0
Ct1
Ct2
Ct3
Ct4
R
AHFM-2005
0.15
0.6
0.6
0.6
1.5
0.5
AHFM-cc
0.2
0.25
0.6
2.5
0.0
0.5
AHFM-NRG
if RePr > 180
0.2
0.053 x ln - 0.27
0.6
2.5
0.0
0.5
Table 4.1: Model constants for the algebraic heat flux model
Note that the nomenclature of the constants used by Shams et al. (2014) will be used in this report.
In the model presented by Kenjereš et al. (2005), an additional coefficient Ct1 has been
introduced as compared to S. Kenjeres (2000) which was by default set to one in earlier
publications.
4.2.1 Equivalence between THFM and ATHFM heat flux modelling
The algebraic expression for the turbulent heat flux can be obtained from the partial differential model
(Eq. 4.4) by simply dropping the unsteady, advection, and diffusion terms to obtain,
(4.19)
where the near wall term in πi and the dissipation rate of the turbulent heat fluxes have been ignored.
By comparison with the algebraic model proposed by Kenjereš et al. (2005) in Eq. 4.17, we note that,
The work of Shams et al. (2014) has shown that Ct1 values are much different from 1.0. For
example, Kenjereš et al. (2005) propose a value of 0.6 which is further reduced by Shams et al. (2014)
for low Prandtl numbers to 0.25 in the so called AHFM-cc model. In the AHFM-NRG model proposed
by them, Ct1 is made into a function of Reynolds and Prandtl numbers with values always below 1.0.
With this in mind, the THFM model is extended by introducing a new coefficient and a new
nomenclature where,
such that a one-to-one equivalence is established between the algebraic model of Kenjereš
et al. (2005) and the differential model of Carteciano & Grötzbach (2003). Due to this
alteration the expression for πi in the transport equations for the turbulent heat fluxes changes
to,
(4.22)
The final set of model coefficients for the THFM model is given in Table 4.2.
uiθ′ tr. equation
θ′2 tr. equation
εθ′ tr. equation
coeff.
value
coeff.
value
coeff.
value
CTD
0.11
CTT
0.13
CDD
0.13
CT0
3.0
CD1
2.2
CT1
0.0
CD2
0.8
CT2
0.33
CP1
1.8
CT3
0.5
CP2
0.72
CT4
0.5
Table 4.2: Set of empirical coefficients for the equivalent THFM
4.3 Algebraic Turbulent Heat Flux Model: θ′2
The ATHFM -θ′2- εθ′ model can be further simplified by calculating the dissipation rate of
the temperature variance εθ′ using the definition of the mechanical to thermal turbulent
time-scale ratio R. The time scale ratio R is either set to a constant value or prescribed
algebraically. The dissipation rate of temperature fluctuation variance can then be directly obtained
as,
(4.23)
This represents a measure of the relative importance of the relaxation timescales of the mechanical and
thermal dissipation. With this simplification, the need to solve a transport equation for εθ′ doesn’t exist.
This model is referred to as the ATHFM -θ′2 model. Although the value and variation of
R in most situations is not known, such an assumption has displayed remarkable success
in a number of thermal flows driven by gravitational effects. Typically, a value of 0.5 is
used.
4.4 WET Model
Further simplification is achieved by determining the turbulent heat flux by applying the WET
(Wealth ≡ Earnings × Time) theory, a syllogism applied by Launder (1988) to turbulent heat
fluxes which lead to the expression: (Value of Second Moment ≡ Production Rate of Second
Moment × Turbulent Time Scale). This, together with the turbulent time k∕ε, yields,
The WET model is supposed to remedy the drawback of simpler variants, in which the heat flux is only
generated by temperature gradients; which is not always the case. For example, the mixed layer formed
close to a heated wall featuring a uniform vertical temperature gradient is not necessarily linked to
turbulence, so the heat flux is actually over-represented in the relative sense. The same is true when
vertical temperature gradients are small: here it is the velocity gradients that cause the wall-to-flow heat
transfer.
The WET model on further simplification yields the GGDH model. The turbulent heat-flux is modelled
in an manner analogous to the turbulent transport term in the Reynolds stress equation, in particular
by reference to Daly & Harlow (1970) model.
(4.25)
This approach is known as the anisotropic eddy-diffusivity model, or the generalized gradient diffusion
hypothesis (GGDH), a definition that points to the fact that heat transfer is driven by an anisotropic
thermal diffusion (Λij),
This approach has the merit to conform to many experimental findings, including the measurements of
turbulent heat transfer in pipes and boundary layer flows by Bremhorst & Bullock (1973), and by many
others. Indeed, these authors demonstrated that turbulent heat flux in the flow direction are two to
three times larger than in the direction normal to the wall while the streamwise temperature gradient is
negligible compared to that normal to the surface.
4.6 Simple Gradient Diffusion Hypothesis (SGDH)
In the context of linear RANS modelling for convective heat transfer, the turbulent heat flux is linked to
the temperature gradient via the expression below where the turbulent stress ui′uj′ is replaced
by the by its trace ui′ui′ equal to 2k, the turbulent kinetic energy. This gives rise to the
isotropic thermal-diffusivity model, known as the Simple Gradient Diffusion Hypothesis
(SGDH).
(4.27)
The turbulent Prandtl number Prt introduced in the above expression is typically set to a constant
value of 0.9. However, two-equation models for thermal diffusivity have been developed in parallel with
the k -ε model of turbulence. This idea was motivated by the fact that turbulent heat diffusion should
also be characterized by a scalar (thermal) time scale that varies in space and time, just like the
turbulent time scale τ = k∕ε. Such models are also referred to as variable turbulent Prandtl number
models.
The model coefficients for the whole chain for models from WET to ATHFM -θ′2-εθ′ is given in
Table 4.3.
Ct0
CΦ
Cε1θ′
Cε3θ′
Cε4θ′
Cε5θ′
Ct1
Ct2
Ct3
Ct4
0.2
0.09
1.3
0.72
2.2
0.8
0.25
0.6
2.5
0.0
Table 4.3: Coefficients for ATHFM
4.7 Variable Turbulent Prandtl Number Model
Within the Reynolds-averaged Navier Stokes model, where the turbulent heat flux is modelled by the
SGDH model, the turbulent Prandtl number Prt is typically assumed to be a constant. This modelling
approach limits the generality of these models because the turbulent Prandtl number can alter the mean
flow prediction. One approach to extend the applicability of the SGDH model is to model the variation
of the turbulent Prandtl number.
The turbulent thermal diffusivity is defined using a composite time scale using both the mechanical
and thermal turbulent time scales given by,
(4.28)
where Cλ is a model constant and fλ is a near-wall damping function for the turbulent thermal
diffusivity applicable for low Reynolds number turbulence models. Using the standard expression for
eddy diffusivity, the turbulent Prandtl number can be expressed as,
(4.29)
In order to close Eq. (4.29) the variance of the temperature fluctuations, and its dissipation rate are
required. If transport equations of these quantities are solved, we obtain the 2-equation variable Prt
turbulent heat flux model.
The equations governing the transport of temperature variance and its dissipation rate are given as,
N. Chidambaram & Kenzakowski (2001)
(4.30)
where the operator denotes the low Reynolds number modifications, ξεθ′ is the near-wall correction
function, and P is the turbulent kinetic energy production term. According to T.P. Sommer &
Lai (1993), different values of Cp1 can be used: a value of 1.8 for boundary layers and 2.0 for internal
flows
(4.32)
The near-wall correction function is given as,
(4.33)
where Pθ′ is the production due to the mean temperature gradient in the streamwise direction to
account for a wall heat-flux boundary, the damping function is defined as,
(4.34)
The near-wall damping function fλ is given as,
(4.35)
Cp1
Cp2
Cp3
Cd1
Cd2
σθ′2
σεθ′
Cλ
A+
C1λ
2.0
0.0
0.72
2.2
0.8
1.0
1.0
0.14
45
0.1
Table 4.4: Coefficients for Variable turbulent Prandtl number model
Chapter 5 Buoyancy–Driven Turbulent Flows
5.1 The Equations of Motion
The buoyancy effects give rise to an additional body-force term fi = -β gi (T -T0) proportional to the
perturbations in the temperature,
this body force per unit volume, applied to a fluid element, has the potential of modifying the turbulent
structure through a production (or destruction) of Reynolds stress at a rate proportional to
(fi′u′j +fj′u′i ), where fi′ is the turbulent perturbation superimposed on fi . The effects of this body
force will change according to the physical processes in question, i.e. through the coupling with the flow
or scalar fields.
Chapter 6 Scale Resolving Strategies: LES and V-LES
6.1 The filtered Navier-Stokes equations (bases of LES)
In Large Eddy Simulation (LES) the motion of the super-grid turbulent eddies is directly
captured whereas the effect of the smaller scale eddies is modelled or represented statistically by
means of simple models, very much the same way as in Reynolds-averaged models (RANS);
i.e. the usual practice is to model the sub-grid stress tensor by an eddy viscosity model.
In terms of computational cost, LES (Sagaut, 2005) lies between RANS and DNS and is
motivated by the limitations of each of these approaches. Since the large-scale unsteady
motions are represented explicitly, LES is more accurate and reliable than RANS. In the
present work, use was made to the Dynamic sub-grid scale model of (Moin et al., 1991);
both in the single-phase and boiling flow case. LES involves the use of a spatial filtering
operation
where the fluctuation of any variable F from its filtered value is denoted by f′ =Fk- Fk. Filter
function G(x-x′) is invariant in time and space, and is localized, obeying the properties:
Applying the filtering operation to the instantaneous Navier-Stokes equations under incompressible
flow conditions leads to the system of filtered transport equations for turbulent convective
flow:
where u is the velocity vector, ρ is the density, π=(-p.I + σ) is the Cauchy stress embedding pressure
and viscous terms. The source terms in the RHS of the momentum equation represents the body force,
Fb, and the convolution-induced, Fc. Further, filtering introduces the SGS stress tensor defined as
τ = ρ(uu-uu), of which the deviatoric part is to be statistically modeled. This way, turbulent scales
larger than the filter width imposed by filter function G(x-x′) are directly solved, whereas the diffusive
effects of the SGS scales are modeled.
6.2 Sub-grid Scale (SGS) Modelling
In turbulent flows, small-scale eddies are known to be simpler to model than the entire spectrum,
since, in this high-wave number zone turbulence is likely to be homogeneous and isotropic
(Kolmogorov, 1942). In other words, the subgrid-scale turbulence is much less problem–dependent than
the resolvable one. Use is generally made of the Eddy Viscosity Concept, linking linearly the SGS eddy
viscosity and thermal diffusivity to the gradients of the filtered velocity and temperature,
respectively:
where Prt is the turbulent Prandtl number, the strain rate tensor Sij and the temperature gradients
are determined from the large-scale turbulence. In the above two relations, the eddy viscosity μt and
thermal diffusivity αθ parameterise locally the non-resolvable dynamic stresses and the heat flux in
terms of the local rate of strain tensor Sij and the temperature gradients, calculated from
the large-scale turbulence. The model constant (Cs) is either fixed or made dependent on
the flow; this latter option is precisely the spirit of the dynamic model, known as DSM
(Moin et al., 1991). A damping function is often introduced for the model constant Cs to
accommodate the asymptotic behavior of near-wall turbulence. Similarly, the same strategy
could be used to close the turbulent SGS heat flux, where the thermal diffusivity could be
determined either based on the resolved thermal-flow field, or alternatively based on the eddy
viscosity (defined dynamically) and a fixed Prt. In the present simulations, we have tested
two variants based on fixed and variable model coefficient Cs, namely the DSM and the
WALE sgs models (Nicoud & Ducros, 1999), the latter has been shown to behave very well
in wall-bounded flows, and to be less dissipative, capable to capture the thin-shear layer
accurately.
6.2.1 The Smagorinsky Model
The Smagorinsky model (Smagorinsky, 1963) is based on the Boussinesq approach, Eqs. (6.6 and
6.7). It is nothing more than a mixing-length theory
where the length scale ℓ is based on the cell size Δ3 = (Δ1Δ2Δ3) rather than on the distance to the
wall, i.e. on the mixing length ℓ0 = κ ⋅ yn . The constant Cs is referred to as the Smagorinsky
constant.
Other options for near-wall treatment are also available. The Deardorff model is given as
(Deardorff, 1970)
(6.9)
where dw is the minimum distance to the wall. In addition to this the Schmidt and Schumann model
(Schmidt & Schumann, 1989) is given as,
where c2 is taken to be c2 = 0.1.
Assuming a direct analogy between momentum and heat flux diffusion by turbulence, one can
determine the thermal diffusivity αθ by reference to the eddy viscosity given by αθ = νt∕Prt , as is
usually done in turbulence modelling. This practice is quite robust in the context of the Smagorinsky
model.
6.2.2 The Dynamic Approach (DSM)
In this approach developed by Germano et al. (1991), A priori adjustment of the Smagorinsky
constant Cs is not needed. Instead, the dynamic model is used to determine this unknown variable Cs
from the information contained in the resolved velocity field.
The main idea consists of introducing a test filter with a larger width than the original one, i.e.
> Δ. This test filter is then applied to the filtered Navier-Stokes equations (the NS equations are
filtered twice), yielding a sub–test–scale stress tensor Tij similar in form to τij that takes the following
form:
By virtue of the Germano identity, the two SGS stress tensors τij and Tij are connected through the
following relation
Assuming now the Smagorinsky functional form to hold and a variable coefficient Cs to be used to
close the deviatoric parts of τij and Tij , we get
where = ( i,j + j,i)∕2 is the rate–of–strain tensor of the test–filtered velocity field. Rearranging the
last three equations results in
A variant of this model (Lilly, 1992) uses a least–squares approach to obtain values for (CsΔ)2,
leading to
Following the procedure employed to derive the dynamic parameter Cs , we can arrive at a relation
for Cθ , the equivalent parameter for the thermal diffusivity, through
We get
where
6.2.3 Turbulent Prandtl Number
The thermal diffusivity can be determined within the Newtonian closure framework using αθ = νt∕Prt ,
but with the turbulent Prandtl number Prt , not αθ , derived via the dynamic procedure of Germano
(Moin et al., 1991). Here as well, information on the resolvable flow and temperature fields serve
to determine the dynamic Prt . The turbulent Prandtl number takes the following form:
Alternatively, a constant turbulent Prandtl number can be assumed, typically (Prt≈ 0.85). The
turbulent Prandtl number can be also calculated based on the model proposed by Kays (1994) given
as,
(6.23)
where Pr = μCp∕λ is the molecular Prandtl number, nu is the kinematic viscosity. As the grid is refined
Prt tends towards large values. This is acceptable since the eddy thermal diffusivity tends to zero.
However, it can be shown that the thermal diffusivity tends to zero much faster than the eddy kinematic
viscosity.
where, gij2 = gikgkj and δij is the Kronecker symbol.
6.3 Variable Turbulence Prandtl Number
Turbulent Prandtl number is given by Kays (Churchill, 2002)
where νt is limited by
6.4 Wall treatment in LES
At high Reynolds number, resolving eddies in the near-wall region would require the
use of a very fine mesh close to the wall. To avoid this, Near-wall treatment is available in
TransAT.
6.4.1 The Werner-Wengle model
The Werner-Wengle wall law (Werner & Wengle, 1991) consists in a two-layer approximation, with
an assumption that a 1∕7th power law holds outside the viscous sub-layer. It can then be
written
(6.30)
with
being the dimensionless wall distance and dimensionless velocity, respectively. ν is the kinematic
viscosity and uτ = is the wall shear stress velocity, with τ the shear at the wall and ρ
the density. The wall shear stress is then computed using the instantaneous velocity in the
flow.
6.5 Very-Large Eddy Simulation Concept (V-LES)
Very Large Eddy Simulation (V-LES) is based on the concept of filtering a larger part of
turbulent fluctuations as compared to LES. This necessitates the use of a more complex
sub–grid modelling strategy. The V-LES used in TransAT based on the use of the k - ϵ model
as a sub–filter model. The filter width can be chosen by the user; it must be larger than
the grid size. Increasing the filter width beyond the largest length scales in the flow, will
lead to a standard RANS simulation, whereas in the limit of a small filter width the model
will tend towards a large eddy simulation (although with a different sub–grid scale model).
The V-LES concept is based on the k -ε model. A filter is applied to this model, so that turbulent
structures smaller than the filter width are not solved. A length-scale limiting function f has been
derived (Johansen et al., 2004) and can be written
(6.33)
where Δ is the filter width and
(6.34)
is a new constant defined by Johansen et al. (2004), with γ ≤ 1 the anisotropic factor and Cμ = 0.09
the k -ε constant. Applying this function to the k -ε model gives the following expression for turbulent
viscosity
(6.35)
Near the wall boundaries, the length-scale limiting function is equal to one, which means that the
standard k - ε model is applied. This enables one to use standard wall-function of RANS models for
V-LES.
Bibliography
Abe, K., Kondoh, T. & Nagano, Y. 1994 A new turbulence model for predicting fluid
flow and heat transfer in separating and reattaching flows–ii. the thermal field calculations. Int.J. Heat Mass Transfer 37, 139–151.
Bremhorst, K. & Bullock, K. 1973 Spectral measurements of turbulent heat and
momentum transfer in fully developed pipe flow. International Journal of Heat and MassTransfer 16 (12), 2141–2154.
Carteciano, L. N. & Grötzbach, G. 2003 Validation of turbulence models in the
computer code flutan for a free hot sodium jet in different buoyancy flow regimes. Tech.Rep. FZKA-6600. Forschungszentrum Karlsruhe GmbH, Karlsruhe, Institut fr Kern- und
Energietechnik.
Churchill, S. W. 2002 A reinterpretation of the turbulent prandtl number. Ind. Eng. Chem.Res. 41, 6393–6401.
Daly, B. & Harlow, F. 1970 Transport equations in turbulence. Phys. Fluids, A 13,
2634–2649.
Deardorff, J. 1970 A numerical study of three-dimensional turbulent channel flow at large
reynolds numbers. J. Fluid Mech. 41, 453–480.
Durbin, P. 1991 Near-wall turbulence closure without damping function. Theoretical &
Comput. Fluid Dynamics 3, 1–13.
Durbin, P. 1993 A reynolds stress model for near-wall turbulence. J. Fluid Mech. 249,
465–498.
Germano, M., Piomelli, U., Moin, P. & Cabot, W. 1991 A dynamic subgrid–scale eddy
viscosity model. Phys. Fluids A 3, 1760–1765.
Hinze, J. O. 1975 Turbulence. New York, USA: MacGraw-Hill.
Johansen, S. T., Wu, J. & Shyy, W. 2004 Filtered-based unsteady RANS computations.
Int. J. of Heat and Fluid Flow 25, 10–21.
Jones, W. & Launder, B. 1972 Prediction of relaminarization with a two-equation
turbulence model. Int. J. Heat Mass Transfer 15, 301–314.
Kays, W. 1994 Turbulent prandtl number - where are we? In The 1992 Max Jacob MemorialLecture. ASME J. Heat Transfer, , vol. 116, pp. 284–295.
Kenjereš, S., Gunarjo, S. & Hanjalić, K. 2005 Contribution to elliptic relaxation
modelling of turbulent natural and mixed convection. International Journal of Heat and FluidFlow 26 (4), 569–586.
Kolmogorov, A. 1942 The equations of turbulent motion in an incompressible fluid. Ivz.Acad. Sci. USSR, Phys. 6, 56–58.
Lam, C. K. G. & Bremhorst, K. 1981 A modified form of the k - ε model for predicting
wall turbulence. Journal of Fluids Engineering 103, 456–460.
Launder, B. 1988 On the computation of convective heat transfer in complex turbulent
flows. ASME J. Heat Transfer 110, 1112–1128.
Launder, B. & Spalding, D. 1974 The numerical computation of turbulent flows. Comput.Meths. Appl. Mech. Eng. 3, 269–289.
Launder, B. E. & Sharma, B. I. 1974 Application of the energy-dissipation model of
turbulence to the calculation of flow near a spinning disc. Letters in Heat and Mass Transfer1, 131–138.
Lilly, D. 1992 A proposed modification of the germano subgrid–scale closure method. Phys.Fluids A 4, 633–635.
Moin, P., Squires, K., Cabot, W. & Lee, S. 1991 A dynamic subgrid-scale model for
compressible turbulence and scalar transport. Phys. Fluids A 3, 2746–2757.
N. Chidambaram, S. D. & Kenzakowski, D. 2001 Scalar variance transport in the
turbulence modeling of propulsive jets. JPP 17.
Nicoud, F. & Ducros, F. 1999 Subgrid-scale stress modelling based on the square of the
velocity gradient tensor. Flow, Turbulence and Combustion 62, 183–200.
Norris, L. & Reynolds, W. 1975 Turbulent channel flow with a moving wavy boundary.
In Rept. No. FM-10. Stanford University, Dept. Mech. Eng.
Rodi, W. 1991 Experience with two–layer models combining the k - ε model with a
one–equation model near the wall. AIAApaper 91–0216.
Rodi, W. & Mansour, N. 1993 Low reynolds k-ε modelling with the aid of direct numerical
simulation. J. Fluid Mech. 250, 509–529.
S. Kenjeres, K. H. 2000 Convective rolls and heat transfer in finite-lenght rayleigh-bnard
convection: A two-dimensional numerical study. Tech. Rep.. Department of Applied Physics,
Thermo-Fluids Section, Delft University of Technology, The Netherlands.
Sagaut, P. 2005 Large Eddy Simulation for Incompressible Flows: An Introduction. Springer.
Sarkar, S., Erlebacher, G., Hussaini, M. Y. & Kreiss, H. O. 1991 The analysis and
modelling of dilatational terms in compressible turbulence. Journal of Fluid Mechanics 227,
473–493.
Schmidt, H. & Schumann, U. 1989 Coherent structure of the convective boundary layer
derived from large-eddy simulation. J. Fluid Mech. 200, 511–562.
Shams, A., Roelofs, F., Baglietto, E., Lardeau, S. & Kenjeres, S. 2014 Assessment
and calibration of an algebraic turbulent heat flux model for low-prandtl fluids. InternationalJournal of Heat and Mass Transfer 79, 589 – 601.
Shih, T.-H., Zhu, J., Liou, W., Chen, K.-H., Liu, N.-S. & Lumley, J. L. 1997 Modeling
of turbulent swirling flows .
Smagorinsky, J. 1963 General circulation experiments with the primitive equations, i, the
basic experiment. Mon. Weather Rev. 91, 99–165.
T.P. Sommer, R. S. & Lai, Y. 1993 Near-wall variable-prandtl-number turbulence model
for compressible flows. AIAA Journal 31 (1), 27–35.
Werner, H. & Wengle, H. 1991 Large-eddy simulation of turbulent flow over and around
a cube in a plate channel. In 8th Symposium on Turbulent Shear Flows. Munich, Germany.
Wilcox, D. 1993 Turbulence Modeling for CFD. DCW Ind., Inc., La Cañada, California.
Yap, C. R. 1987 Turbulent heat and momentum transfer in recirculating and impinging
flows. Doctoral thesis, Univ. Manchester, United Kingdom.