16.08.2013 Views

Scalar Transport - Turbulence Mechanics/CFD Group

Scalar Transport - Turbulence Mechanics/CFD Group

Scalar Transport - Turbulence Mechanics/CFD Group

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

School of Mechanical Aerospace and Civil Engineering<br />

TPFE MSc Advanced <strong>Turbulence</strong> Modelling<br />

<strong>Scalar</strong> <strong>Transport</strong><br />

T. J. Craft<br />

George Begg Building, C41<br />

Reading:<br />

S. Pope, Turbulent Flows<br />

D. Wilcox, <strong>Turbulence</strong> Modelling for <strong>CFD</strong><br />

Closure Strategies for Turbulent and Transitional<br />

Flows, (Eds. B.E. Launder, N.D. Sandham)<br />

Notes: Blackboard and <strong>CFD</strong>/TM web server:<br />

http://cfd.mace.manchester.ac.uk/tmcfd<br />

- People - T. Craft - Online Teaching Material<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 1 / 18<br />

◮ Equation (1) can be averaged in the same way as the momentum<br />

equations. First, we split Φ into mean and fluctuating parts:<br />

Φ(x,t) = Φ(x) + φ(x,t) (2)<br />

◮ Substituting into equation (1) and averaging leads to an equation of the<br />

form<br />

∂Φ ∂ <br />

+ UjΦ<br />

∂t ∂xj = ∂<br />

<br />

λ<br />

∂xj ∂Φ<br />

<br />

− u<br />

∂x<br />

jφ +<br />

j<br />

SΦ<br />

(3)<br />

where SΦ represents the averaged source terms from the original<br />

instantaneous equation.<br />

◮ Comparing with the Reynolds averaged Navier-Stokes equations, in this<br />

case the additional terms involve the turbulent scalar fluxes, u jφ.<br />

◮ In order to close the equation, we need to devise models for these<br />

turbulent scalar fluxes.<br />

◮ For simplicity, we consider the problem of modelling turbulent heat fluxes,<br />

uiθ, although the principles are easily extended to other scalars.<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 3 / 18<br />

Introduction<br />

◮ In many <strong>CFD</strong> applications we may be interested in predicting the<br />

behaviour of scalar quantities such as enthalpy, temperature, mass<br />

fraction of a chemical species, etc.<br />

◮ Consider a property Φ transported by the flow. We assume that its<br />

evolution can be described by the equation<br />

∂ <br />

Φ<br />

∂t<br />

∂<br />

<br />

+ Uj Φ<br />

=<br />

∂xj ∂<br />

∂xj λ ∂ <br />

Φ<br />

+<br />

∂xj SΦ<br />

where SΦ represents any source or sink term that may be present.<br />

◮ If Φ is enthalpy or temperature, the flow is low speed and the effects of<br />

viscous heating are negligible, then SΦ can often be ignored.<br />

◮ If Φ represents the mass fraction of a species being consumed or<br />

produced by chemical reaction, then SΦ may be of great importance and<br />

has to be modelled from the details of the reaction process.<br />

◮ If Φ represents temperature, then λ is simply ν/Pr where Pr is the<br />

molecular Prandtl number of the fluid.<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 2 / 18<br />

Eddy-Diffusivity Modelling<br />

◮ A linear eddy-viscosity model approximates the turbulent stresses by:<br />

<br />

∂Ui<br />

uiuj = (2/3)kδij − νt +<br />

∂xj ∂U <br />

j<br />

(4)<br />

∂xi ◮ An obvious extension of this is to introduce an eddy-diffusivity for the<br />

scalar fluxes, related to the turbulent viscosity:<br />

u iθ = − νt<br />

σt<br />

∂Θ<br />

∂x i<br />

◮ The total flux terms in the mean temperature transport equation then<br />

become<br />

<br />

∂ ν νt ∂Θ<br />

+<br />

Pr σt<br />

∂x k<br />

◮ The turbulent Prandtl number, σt, is often assumed to be constant.<br />

◮ In a near-wall flow, σt = 0.9 is usually adopted. However, in free flows a<br />

lower value (around 0.7) is more appropriate.<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 4 / 18<br />

∂x k<br />

(1)<br />

(5)


<strong>Scalar</strong> Fluxes in Simple Shear<br />

◮ An exact equation can be derived for u iθ, which shows it has a<br />

generation term of the form<br />

P iθ = −u iu k<br />

∂Θ<br />

− u<br />

∂x<br />

kθ<br />

k<br />

∂Ui ∂xk ◮ In simple shear flow U = U(y), Θ = Θ(y), this gives:<br />

P1θ = −uv dΘ<br />

dy<br />

dΘ<br />

P2θ = −v 2<br />

dy<br />

− vθ dU<br />

dy<br />

◮ This suggests both the wall-normal and wall-parallel scalar fluxes will be<br />

non-zero.<br />

◮ Measurements and DNS data show that generally |uθ | > |vθ |,<br />

particularly in the near-wall region.<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 5 / 18<br />

◮ However, the eddy diffusivity model gives:<br />

vθ = − νt dΘ<br />

σt dy<br />

U(y)<br />

and uθ = 0<br />

ie. zero turbulent scalar flux in the streamwise direction.<br />

◮ Making the usual boundary layer approximations, the (steady) mean<br />

scalar transport equation becomes<br />

U ∂Θ<br />

<br />

∂Θ ∂<br />

+ V = α<br />

∂x ∂y ∂y<br />

∂Θ<br />

<br />

− vθ<br />

∂y<br />

◮ Consequently, if streamwise gradients are small compared to<br />

cross-stream ones, the above error in uθ may not have serious<br />

consequences.<br />

◮ However, the error may be more serious in buoyancy-influenced or other<br />

complex flows.<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 7 / 18<br />

y<br />

x<br />

Θ(y)<br />

Homogeneous Shear Flow<br />

y<br />

x<br />

U(y)<br />

Θ(y)<br />

Expts: Tavoularis & Corrsin (1981)<br />

y<br />

x<br />

Chevray & Tutu (1978)<br />

Expts:<br />

q<br />

U(y)<br />

T(y)<br />

q<br />

Heated Plane Channel<br />

y x<br />

DNS: Horiuti (1992)<br />

Heated Axisymmetric Jet<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 6 / 18<br />

Generalized Gradient Diffusion Modelling<br />

◮ A somewhat better model for the scalar fluxes arises from the Daly &<br />

Harlow (1970) Generalised Gradient Diffusion Hypothesis (GGDH):<br />

◮ In a simple shear flow this gives<br />

k dΘ<br />

vθ = −cθ v 2<br />

ε dy<br />

k<br />

uiθ = −cθ<br />

ε u ∂Θ<br />

iuj ∂xj k dΘ<br />

and uθ = −cθ uv<br />

ε dy<br />

◮ This does at least give a non-zero uθ, and also better reflects the<br />

dependence of vθ on v 2 , seen in the generation terms.<br />

◮ If a stress transport model or non-linear EVM is used which gives good<br />

stress anisotropy levels, the above form can be used as is (typically with<br />

a coefficient cθ around 0.3).<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 8 / 18


◮ With an EVM (which will not generally provide good normal stress<br />

predictions), the GGDH can still be used with some modification.<br />

◮ Ince & Launder (1989) used this model in conjunction with a linear<br />

eddy-viscosity model for the stresses, taking cθ = (3/2)(cμ/σt).<br />

◮ In a shear flow (where the model returns v 2 = (2/3)k) this gives<br />

k dΘ cμ k<br />

vθ = −cθ v 2 = −<br />

ε dy σt<br />

2 dΘ<br />

ε dy<br />

k dΘ<br />

uθ = −cθ uv<br />

ε dy<br />

which gives vθ as in the eddy-diffusivity model, but should give a better<br />

approximation of uθ.<br />

◮ More complex algebraic scalar flux models have been proposed, some<br />

developed along similar routes to the non-linear eddy-viscosity models<br />

examined earlier for the Reynolds stresses.<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 9 / 18<br />

◮ Density fluctuations are often expressed in terms of temperature<br />

fluctuations, by introducing the thermal expansion coefficient:<br />

β = − 1 ∂ρ<br />

ρ ∂Θ<br />

◮ The buoyancy source term in the k equation then becomes<br />

(9)<br />

G k = −β g i u iθ (10)<br />

◮ Note that the buoyant generation of k depends on the turbulent scalar<br />

fluxes.<br />

◮ Depending on the sign of vθ, buoyancy can either enhance or reduce k<br />

levels:<br />

◮ In a stably stratified layer (with dΘ/dy > 0) we would typically have<br />

vθ < 0 and hence (with g2 < 0), G k is negative.<br />

◮ In an unstable layer, dΘ/dy < 0, and hence vθ > 0 and G k is<br />

positive.<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 11 / 18<br />

Buoyancy Effects<br />

◮ One important application where temperature (or concentration)<br />

differences must be properly accounted for is in buoyancy-affected flows.<br />

◮ In that case, there is an additional source term in the momentum<br />

equation:<br />

D(ρUi)<br />

Dt = ... + ρg i (6)<br />

◮ This results in an additional term in the transport equation for the<br />

fluctuating velocity:<br />

Du i<br />

Dt = ... + ρ′ g i/ρ (7)<br />

where ρ ′ is the fluctuating density.<br />

◮ Consequently, one gets an additional generation term in the turbulent<br />

kinetic energy equation:<br />

G k = (1/ρ)g i u iρ ′ (8)<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 10 / 18<br />

Buoyancy in Stress <strong>Transport</strong> Models<br />

◮ The buoyancy-related term appearing in the transport equation for u i<br />

leads to an additional generation term in the stress transport equations:<br />

where<br />

Du iu j<br />

Dt = P ij + G ij + φ ij − ε ij + d ij (11)<br />

G ij = −β g i u jθ − β g j u iθ (12)<br />

◮ When modelling the pressure-strain redistribution process, a contribution<br />

due to buoyancy should also be included:<br />

φ ij = φ ij1 + φ ij2 + φ ij3<br />

◮ Within the framework of the modelling adopted earlier, φ ij3 is<br />

approximated in a similar manner to φ ij2:<br />

(13)<br />

φ ij3 = −c3(G ij − (1/3)G kkδ ij) (14)<br />

◮ φ ij3 is thus assumed to redistribute the buoyancy generation.<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 12 / 18


◮ In some instances it may be sufficient to model the scalar fluxes u iθ<br />

using a GGDH approach (or similar) as described earlier.<br />

◮ However, in strongly buoyant flows this may not be adequate, since the<br />

scalar fluxes are themselves affected by buoyancy.<br />

◮ Some extended algebraic heat flux models have been proposed,<br />

incorporating some buoyancy effects (eg. Hanjalić et al, 1996).<br />

◮ Another option is to adopt a full second-moment closure, solving<br />

transport equations for the scalar fluxes also.<br />

◮ <strong>Scalar</strong> flux transport equations can be derived in a similar manner to<br />

those for the Reynolds stresses. The result can be written in the form<br />

Du iθ<br />

Dt = Piθ + Giθ + φiθ − εiθ + diθ (15)<br />

◮ The generation terms P iθ and G iθ are exact:<br />

P iθ = −u iu k<br />

∂Θ<br />

− u<br />

∂x<br />

kθ<br />

k<br />

∂Ui ∂xk (16)<br />

G iθ = −β g i θ 2 (17)<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 13 / 18<br />

Negatively-Buoyant Jet<br />

◮ Axisymmetric downward<br />

directed buoyant jet.<br />

◮ Experiment of Cresswell et al<br />

(1989).<br />

◮ Velocity and shear stress<br />

profiles one diameter<br />

downstream of jet discharge.<br />

—— TCL RSM; – – Basic RSM<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 15 / 18<br />

◮ Other terms in the transport equation have to be modelled.<br />

◮ Similar assumptions are typically made for these as for the corresponding<br />

terms in the stress transport equations. However, the details will not be<br />

covered in this course.<br />

◮ The scalar flux buoyancy generation depends on the scalar variance, θ 2 .<br />

◮ This can be obtained by solving its transport equation, of the form<br />

where the generation rate is given by<br />

Dθ 2<br />

Dt = Pθ − 2εθ + diffusion (18)<br />

Pθ = −2u iθ ∂Θ<br />

◮ The dissipation rate εθ is usually either modelled algebraically by<br />

assuming that thermal and dynamic timescales are related (eg.<br />

R = (k/ε)(2εθ /θ 2 ) being constant), or obtained from its own modelled<br />

transport equation.<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 14 / 18<br />

Opposed/Buoyant Wall Jet<br />

∂xi<br />

◮ Isothermal and buoyant cases studied.<br />

◮ LES data from Addad et al (2004).<br />

◮ Requires a good outer flow and near-wall<br />

modelling for accuracy.<br />

◮ Vertical<br />

velocity<br />

contours.<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 16 / 18


Stably Stratified Mixing Layer<br />

◮ Studied experimentally by<br />

Uittenbogaard (1998).<br />

◮ U1 = 0.5m/s, ρ1 = 1015kg/m 3<br />

◮ U2 = 0.3m/s, ρ2 = 1030kg/m 3<br />

◮ Linear k-ε scheme overpredicts<br />

turbulence levels and hence mixing.<br />

◮ Second-moment closures do better at<br />

capturing buoyancy effects on scalar<br />

fluxes and hence reduce mixing.<br />

◮ In this case, as turbulence levels<br />

decrease, modelling of diffusion<br />

becomes more influential.<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 17 / 18<br />

References<br />

◮ Addad, Y., Benhamadouche, S., Laurence, D., (2004) The negatively buoyant jet: LES<br />

results, Int. J. Heat and Fluid Flow, vol. 25, pp. 795-808.<br />

◮ Chevray, R., Tutu, N.K., (1978) Intermittency and preferential transport of heat in a round jet,<br />

J. Fluid Mech., vol. 88, p. 133.<br />

◮ Cresswell, R., Haroutunian, V., Ince, N.Z., Launder, B.E., Szczepura, R.T., (1989)<br />

Measurement and modelling of buoyancy-modified elliptic turbulent shear flows, Proc. 7th<br />

Turbulent Shear Flows Symposium, Stanford University.<br />

◮ Daly, B.J., Harlow, F.H., (1970) <strong>Transport</strong> equations in turbulence, Phys. Fluids, vol. 13, pp.<br />

2634-2649.<br />

◮ Hanjalić, K., Kenjeres, S., Durst, F., (1996) Natural convection in partitioned two-dimensional<br />

enclosures at higher Rayleigh numbers, Int. J. Heat Mass Transfer, vol. 39, pp. 1407-1427.<br />

◮ Horiuti, K., (1992) Assessment of two-equation models of turbulent passive-scalar diffusion<br />

in channel flow, J. Fluid Mech., vol. 238, pp. 405-433.<br />

◮ Ince, N.Z., Launder, B.E., (1989) On the computation of buoyancy-driven turbulent flow in<br />

rectangular enclosures, Int. J. Heat Fluid Flow, vol. 10, pp. 110-117.<br />

◮ Tavoularis, S., Corrsin, S., (1981) Experiments in nearly homogeneous turbulent shear flow<br />

with a uniform mean temperature gradient. Part 1., J. Fluid Mech., vol. 104, pp. 311-347.<br />

◮ Uittenbogaard, R.E., (1998) Measurement of turbulence fluxes in a steady, stratified, mixing<br />

layer, Proc. 3rd Int. Symposium on Refined Flow Modelling and <strong>Turbulence</strong> Measurements,<br />

Tokyo.<br />

<strong>Scalar</strong> <strong>Transport</strong> 2011/12 18 / 18

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!