Вы находитесь на странице: 1из 7

REVIEW

pubs.acs.org/IECR

On the Gradient Diffusion Hypothesis and Passive Scalar Transport in


Turbulent Flows
Daniel P. Combest, Palghat A. Ramachandran, and Milorad P. Dudukovic*
The Chemical Reaction Engineering Laboratory, Department of Energy, Environmental, and Chemical Engineering,
Washington University, St. Louis, Missouri 63130, United States

ABSTRACT: A discussion of modeling passive scalar transport in turbulent flows is given. Several methods employed to close the scalar-
flux term Æuuφuæ that arises during Reynolds averaging are provided. Alternatives and improvements to the gradient diffusion hypotheses
are addressed, most notably, the need for an alternative to the global constant turbulent Schmidt and Prandtl numbers. The reader is
given a brief history covering methods used to predict turbulent Schmidt and Prandtl numbers, along with recommendations for future
research, based partially on studies by Professor Stuart Churchill. More detailed formulations of turbulent Schmidt or Prandtl numbers
will enable better approximations of the influence of turbulence in models of passive scalar flows using the gradient diffusion hypothesis.

1. INTRODUCTION 2. CONSERVED SCALARS IN TURBULENT FLOWS


The development of accurate mathematical descriptions and A scalar is defined as a rank-zero tensor representing a simple
computationally efficient numerical methods representing react- physical quantity (e.g., volume fraction, mass, temperature, or
ing turbulent flow is paramount to the future of reactor modeling species concentration) in a system. Subsequently, transport
in chemical reaction engineering (CRE). As computational capa- equations are used to describe the conservation of a scalar within
bilities increase with the improvement of numerical algorithms a system to represent inflow, outflow, generation, and accumula-
and computing hardware, CRE moves toward more-detailed tion of a scalar quantity. If scalar transport does not interact with
multiscale models of chemical reactors. The methods and com- concurrent transport processes in the system, it is appropriately
plexity of such technology aim to maximize description with the named a passive scalar. Conversely, if the scalar affects other
constraint of computational cost and practicality. This begs the simultaneous transport processes, it is referred to as an active
question: How can an existing method be employed more scalar (e.g., thermal interactions with momentum through a
efficiently to produce more accurate and useful results? This
buoyancy term in the momentum conservation equation). Scalar
paper strives to address this question, with regard to the modeling
transport (passive or active) is present throughout many scien-
of passive scalar transport in a turbulent flow using a Reynolds-
averaging approach, the use of the gradient-diffusion hypothesis to tific fields, from basic science in the study of transport phenom-
close the scalar-flux term, and the turbulent Schmidt (Sct = νt /Dt ) ena to industrial engineering processes. In the case of a scalar in
and Prandtl (Prt = νt /rt) number-based approximations of the a flowing fluid, turbulence is often encountered and requires
turbulent mass and thermal diffusivity. additional consideration.
The following discussion in this paper is mainly aimed at the Although our understanding of turbulent flow has grown
graduate student or researcher, providing a brief background, tremendously over the last century, fundamental knowledge of
building up to the suggested improvement of Sct and Prt estima- the mechanism of conserved scalar transport in a turbulent flow is
tion techniques. Section 2 defines the scalar and gives a discus- still a growing field of research. Recent reviews by Dimotakis,1
sion on scalar mixing in turbulent flows. Section 3 reviews tur- Warhaft,2 and Tominaga and Stathopoulos3 provide valuable
bulence modeling using the Reynolds-averaged NavierStokes background covering the experimental, theoretical, and modeling
(RANS) equations, enabling the reader to become familiar with work concerning scalar transport during the 20th century.
the turbulent closure problem. Section 4 introduces the transport Dimotakis reviewed turbulent mixing, with a thorough discussion
of passive scalars, while several methods of closing the scalar-flux into what has been categorized as level-1, level-2, and level-3
term are addressed in section 5. Section 6 discusses the meaning mixing,1 where
of the Sct and Prt, with a review of the numerical determination • level-1 mixing involves passive scalars, such that the combi-
of Sct and Prt. Finally, section 7 discusses a reinterpretation of nation of a fluid and tracer or fluids of similar properties
the Prt by Professor Stuart Churchill and a look toward the near result in no effect on overall flow dynamics.
future of Sct and Prt prediction. Overall, the reader should be left
with an understanding of the typical usage of Sct and Prt in
Special Issue: Churchill Issue
computational fluid dynamics (CFD) simulations. Conclusions
and recommendations for future research that fully leverages Sct Received: January 10, 2011
and Prt to more accurately predict passive scalar transport in Accepted: May 4, 2011
turbulent flows using Reynolds-averaged models are given in the Revised: April 12, 2011
final section. Published: May 04, 2011

r 2011 American Chemical Society 8817 dx.doi.org/10.1021/ie200055s | Ind. Eng. Chem. Res. 2011, 50, 8817–8823
Industrial & Engineering Chemistry Research REVIEW

• level-2 mixing is indicated by the interaction of two fluids 4 is referred to as the unsteady Reynolds-averaged Navier
causing a change in flow dynamics (e.g., RayleighTaylor Stokes equation (or URANS). For steady-state turbulent flows,
instability flows). the time derivative in eq 4 is zero and the remaining equation is
• level-3 mixing is categorized by mixing that produces referred to as the Reynolds-averaged NavierStokes (or RANS)
changes in the overall fluid-intensive properties (density equation. Equation 4 presents a Reynolds stress term, F Æuuu
i uæ
j as a
or composition), resulting in a change in flow dynamics cross correlation of uiu and uju, leaving more unknowns than
(e.g., buoyancy-driven flow). equations. This is referred to as the turbulence closure problem.
Dimotakis noted that studies on level-2 and level-3 mixing are To deal with the issue of closure, the problem can be ap-
very much open research topics, while level-1 mixing research has proached in several different ways, three of which are presented in
been limited to canonical flows (e.g., pipe flow, free shear layers, this discussion. The first method is to approximate the unclosed
and jets) and relations to empirical data.1 The conclusions drawn Æuiuujuæ term in the RANS and URANS equations with the Boussinesq
by Dimotakis1 end on the hope for experimental studies to pro- eddy viscosity hypothesis,7 given as8
duce more-detailed data needed to develop large eddy simula- !  
tion (LES) subgrid scale (SGS) models based on observations of 0 0 DÆUi æ DÆUj æ 2 DÆUk æ
 FÆui uj æ ¼ μt þ  δij μt þ Fk
scalar mixing, rather than on low-order statistics. Warhaft focused Dxj Dxi 3 Dxk
on the turbulent passive scalar, touching deeply on passive scalar
anisotropy and the direct connection of turbulent length scales, ð5Þ
rather than the cascading mechanisms of turbulent interactions.2 where k is the turbulent kinetic energy (k = uuu
i u/2),
j δij is the
Mainly giving an experimental perspective, cornerstones of tur- Kronecker delta,4 and μt is the turbulent viscosity scalar
bulence theory are addressed in an effort to understand inter- determined by a turbulent viscosity model. Over the last
mittency. However, Warhaft’s discussion,2 which is fundamentally century, numerous turbulent viscosity models have been de-
important in understanding turbulence, is beyond the general vised, including kε, kω, SpalartAllmaras, and RNG kε.
remarks of this review and left to the motivated reader. Lastly, the These models, along with numerous other formulations, are
work by Tominaga and Stathopoulos offers a discussion of the covered in detail in many books on fluid dynamics and
Reynolds-averaged approach to modeling passive scalars, with an computational continuum mechanics.8,9
emphasis on gas-phase urban and building diffusion problems.3 The second method for determining the unclosed stress F
The overall conclusion from Tominaga and Stathopoulos3 is that Æuiuujuæ involves modeling the Reynolds stresses with transport
there is a need for optimal global Sct numbers based on dominant equations. The model derivation for the Reynolds stresses trans-
effects, instead of the generally accepted values of the turbulent port equations is thoroughly covered in many books on compu-
Schmidt (Sct) number in the range of 0.70.9. The implications of tational fluid dynamics.8,10 The final equation11 of the transport
a turbulent Schmidt number are provided later in this discussion. of Reynolds stresses is expressed as
The cited reviews provide an appropriate introduction into the
topic of passive scalar transport in turbulent flows, but leave room 0
DÆui uj æ
0 0 0
DÆui uj æ
0 0
DÆui uj uk æ
0
0 0
for further improvement on the subject in terms of Reynolds- þ ÆUk æ þ ¼ Pij þ Πij þνrÆui uj æ  εij
averaged modeling approaches. Dt Dxk Dxk
ð6Þ
3. THE REYNOLDS-AVERAGED NAVIERSTOKES
with the production term being defined as
EQUATION
In the Reynolds-averaged NavierStokes (RANS) equations, 0 0 DÆUj æ 0 0 DÆUi æ
Pij ¼  Æui uk æ þ Æuj uk æ ð7Þ
turbulent momentum conservation is described using the Navier Dxk Dxk
Stokes equations coupled with a perturbation in the velocity and
pressure variables.4 The approach uses a time- and space-dependent the velocitypressure gradient term being defined as
velocity variable U(x,t) decomposed into an ensemble-averaged  
1 0 Dp0 0 Dp
0
velocity ÆUæ and a fluctuating velocity uu component, where Πij ¼  ui þ uj ð8Þ
F Dxj Dxi
0
Uðx, tÞ ¼ ÆUðx, tÞæ þ u ðx, tÞ ð1Þ
and the turbulent dissipation term
Similarly, the pressure can be represented in terms of the Reynolds  0 0
decomposition of the original variable p(x,t), so that Dui Duj
εij ¼ 2ν ð9Þ
pðx, tÞ ¼ Æpðx, tÞæ þ p0 ðx, tÞ ð2Þ Dxk Dxk

By substituting eqs 1 and 2 into the NavierStokes equations, The triple-correlation term Æuiuujuuku æ is closed with an assumption
the continuity equation becomes similar to the Boussinesq eddy viscosity hypothesis.
An additional approach used to close the Reynolds stress term
r 3 ðÆUæ þ u0 Þ ¼ 0 ð3Þ involves algebraic Reynolds stress models (ASMs). These types
of models assume a nonlinear constitutive model to relate the
and momentum conservation is described by
Reynolds stresses and the rate of mean strain rather than addi-
DÆUi æ DÆUi æ 1 D 0 0 1 DÆpæ tional transport equations for each of the Reynolds stresses.8
þ ÆUj æ ¼ νr2 ÆUi æ  Æu u æ  ð4Þ These relationships are held to physical and mathematical con-
Dt Dxj F Dxj i j F Dxj
straints including Galilean invariance and “realizability” to ensure
In eqs 3 and 4, ÆUæ and uu are solenoidal vector fields,5,6 ν is the physically meaningful results. Lastly, the basis of ASMs require
molecular kinematic viscosity, and F is the fluid density. Equation that adjustments be made to adequately predict turbulent flow
8818 dx.doi.org/10.1021/ie200055s |Ind. Eng. Chem. Res. 2011, 50, 8817–8823
Industrial & Engineering Chemistry Research REVIEW

and generally do not perform as well as a full Reynolds- model that is similar to the eddy viscosity model (eq 5), an
stress model. algebraic moment (AM) model, or a scalar-flux transport model
The RANS equation and subsequent closure relations provide (SFM). Each of these approaches differ in formulation, as well as
valuable insight that has been applied to Reynolds-averaged complexity of transport equations, and has strengths and weak-
passive scalar transport discussed in later sections. What is impor- nesses that will be covered in this discussion. The full derivations
tant to note is that gradient diffusion hypothesis approximations and applications of each of these methods are available in the
are used extensively throughout RANS model derivation in the cited literature.
Boussinesq eddy viscosity hypothesis (eq 5) and throughout 5.1. Gradient Diffusion Models. By far, the simplest method
ASMs. What is also integral to the discussion has been addressed to account for ÆuiuφRu æ is to use a gradient diffusion hypothesis
by Tannehill, Anderson, and Pletcher,8 stating that transport (GDH), in which
equations can be written to close almost any term, but none of 0 νt
them can be solved exactly. This idea, along with closure tech- Æu0 φR æ ¼  Dt rÆφR æ ¼  rÆφR æ ð15Þ
Sct
niques for the Reynolds-stress term Æuiuujuæ, are applied and dis-
cussed in the next section concerning the Reynolds-averaged with Sct being the turbulent Schmidt (or turbulent Prandtl)
passive scalar transport. number and Dt being the turbulent mass diffusivity. The gradient
diffusion hypothesis assumes isotropic turbulence for simplicity.
4. THE REYNOLDS-AVERAGED PASSIVE SCALAR As a result, the GDH is known to inaccurately predict turbulent
EQUATION effects in cases where the scalar flux is not aligned with the mean
Inasmuch the RANS equation requires closure to account for scalar gradient (i.e., highly anisotropic flows). The benefit of the
the effect of chaotic turbulent fluctuations, so does the transport GDH closure is that there are no additional transport equations,
of a passive scalar in a turbulent flow field.5,11 The conservation making it relatively simple to implement numerically.
equation describing the scalar transport of species R is 5.2. Algebraic Models. A more rigorous approach than the
GDH involves determining an anisotropic turbulent diffusivity
DφR tensor Dtij, which was first introduced by Batchelor12 as
þ r 3 ðUφR Þ ¼ r 3 DR rφR þ SR ðφÞ ð10Þ
Dt
0 0 DÆφR æ
where SR(φ) is the chemical source term (i.e., rate of production) Æui φR æ ¼  Dtij ð16Þ
Dxj
for species R. Using Reynolds decomposition,4 a scalar quantity
φR is decomposed to an ensemble-averaged value ÆφRæ and a with the difficulty being observed in directly determining the
fluctuating component φRu as components of Dtij. Younis13 noted that the simplest rational
0 algebraic model (AM) was developed by Daly and Harlow,14 by
φR ðx, tÞ ¼ ÆφR ðx, tÞæ þ φR ðx, tÞ ð11Þ setting Dtij directly proportional to the Reynolds stresses, with
Substituting eq 11 into the scalar transport equation for species  
0 0 k 0 0 DÆφ æ
R in eq 10, we are left with Æui φR æ ¼  Cθ Æui uj æ R ð17Þ
ε Dxj
DÆφR æ
þ r 3 ÆUφR æ ¼ r 3 DR rÆφR æ þ ÆSR ðφÞæ ð12Þ where Cθ is set as a positive constant. This formulation over-
Dt
comes one of the shortcomings of GDH due to misalignment of
where the second term on the left-hand side of eq 12 requires scalar flux and mean scalar gradient but, as Younis13 comments,
further decomposition to eq 17 predicts the wrong magnitude of the scalar flux in the
r 3 ÆUφR æ ¼
0
r 3 ÆðÆUæ þ u0 ÞðÆφR æ þ φR Þæ direction normal to the mean scalar gradient. The conclusion by
0 ð13Þ Younis was drawn through comparison of the ratio of streamwise
¼ r 3 ðÆUæÆφR æ þ Æu0 φR æÞ to cross-stream heat fluxes in fully developed channel flow with
heated walls, noting the slight differences of the DalyHarlow
Back substitution of eq 13 into eq 12 results in the unsteady model (eq 17), compared to the DNS results by Kim.15 A similar
Reynolds-averaged passive scalar equation model that was discussed by Fox11 reveals a slight variation,
DÆφR æ 0
where
þ r 3 ðÆUæÆφR æÞ ¼ r 3 DR rÆφR æ  r 3 Æu0 φR æ þ ÆSR ðφÞæ
Dt 0 0 k 0 0 DÆφR æ
ð14Þ Æui φR æ ¼  Æu u æ ð18Þ
Sct ε i j Dxj
As it was seen in the RANS equation, the unclosed ÆuuφRu æ scalar-
which is used in conjunction with the kε or Reynolds-stress
flux term in eq 14 involves fluctuating components that must be
models. Equation 18 correspondingly overcomes the flaw of
reconciled. In the case of production (i.e., nonzero chemical source
scalar flux and mean gradient misalignment of the GDH, because
SR(φ)), additional numerical treatment is required to close the
of its use of the anisotropic Æuuu
i uæ
j term. Both eqs 17 and 18 draw
ensemble-averaged chemical source term ÆSR(φ)æ. Further discus-
criticism11,13 for their lack of accuracy in predicting the scalar
sion of the chemical source closure can be found in works by Pope5
flux; yet, they both imply a strong influence of Reynolds stress on
and Fox.11 The closure of the ÆuuφuRæ term in eq 14 is discussed in
simple scalar-flux models. This is a more agreeable description of
the next section.
physical phenomena in turbulent flows than those relying on the
simplifying assumptions of turbulent isotropy. The theme of
5. THE CLOSURE OF THE SCALAR-FLUX TERM ÆuuφRu æ Reynolds-stress influence on scalar flux is recurrent in commen-
The closure of the ÆuuφRu æ term in eq 14 is generally accom- taries by Churchill,16 related directly to Prt, and the discussions in
plished with either a gradient-diffusion hypothesis (GDH) classic literature by Levich17 on diffusion rates in turbulent
8819 dx.doi.org/10.1021/ie200055s |Ind. Eng. Chem. Res. 2011, 50, 8817–8823
Industrial & Engineering Chemistry Research REVIEW

boundary layers. Not surprisingly, more extensive models used problem-dependent. In all cases, a sensitivity study of a particular
to close the scalar-flux yield more detailed descriptions of tur- variable correlated to Sct should be performed.28
bulent effects on scalar transport. Hence, one may more accu- In contrast to the use of a global Sct and Prt to lump the com-
rately resolve scalar flux by utilizing additional transport equa- plex relationship of turbulent fluctuations and turbulent mixing
tions, which are discussed in the next section. through values of νt and Dt or Rt, very little work has been done
5.3. Scalar-Flux Models. Similar to the Reynolds-stress over the last few decades to elucidate the factors influencing Sct
equations, in a scalar-flux model (SFM) approach, the ÆuuφRu æ is and Prt. As addressed in a 1974 article by A. J. Reynolds,29 the
treated directly11,13 with the transport equation: previous 25 years of predictions of the turbulent Schmidt number
(Sct = νt/Dt) and Prandtl number (Prt = νt/Rt) were reviewed.
0 0 0 0  
DÆui φR æ DÆu φ æ D 0 0 0 1 0 The discussion was restricted to temperature and concentration
þ ÆUj æ i R ¼ Jij  Æuj ui φR æ  Æp0 φR æδij ranges that do not significantly interact with the flow (i.e., passive
Dt Dxj Dxj F
scalar flows and level-1 mixing), and assumed an almost-perfect
þ Pi þ Ri  εi ð19Þ analogy between Prt and Sct numbers, with minor exception in
gases. This exception is based on the differences between actual
where Jij is the molecular diffusion component (which is often and assumed enthalpy fluxes seen in the flows, where they are
neglected, since molecular diffusion will be small compared to tur- negligible and within measured experimental variability. Rey-
bulent diffusion effects in higher Reynolds number flows); Pi is the nolds noted that Prt and Sct depended on the corresponding
closed production term; Ri is the unclosed pressure-scalar-gradient molecular value of Pr and Sc, the position within the flow (i.e.,
term; and εi is the scalar-flux dissipation term (which is often distance from the wall), and the local turbulent intensity of the
neglected if small-scale isotropy is assumed). The remaining flow in question. In general, Reynolds summarized these ob-
ÆujuuiuφRu æ and ÆpuφRu æ flux terms are generally closed by relations servations into a single general formula:
similar to the gradient-diffusion hypothesis.11 The triple correla- "  n #
tion term Æuuu j uφ
i uRæ in eq 19 should be of much lower magnitude, νt
Sct ¼ C1 exp  C2 Scm ð20Þ
compared to the double correlation term ÆuiuφRu æ in eq 14, making ν
the use of GDH to describe ÆujuuiuφRu æ, which is an acceptable
approximation of less-dominant phenomena in the SFM. The SFM where C1, C2, m, and n are all positive constants. In eq 20, the position
is one of the most-detailed Reynolds-averaged models for scalar-flux within the flow is taken into account implicitly through the νt term,
transport, and it is the most computationally intensive method, which is spatially variable and highly influenced by the wall and
compared to the GDH and AM presented in this discussion. turbulent boundary layer. Equation 20 is consistent with the observa-
tion that Sct decreases as νt/ν increases, indicating greater turbulent
mixing in more-turbulent regions. For νt/ν ratios approaching zero,
6. THE TURBULENT SCHMIDT AND PRANDTL Sct approaches C1. This is inconsistent with the definition of Sct = νt/
NUMBERS Dt, because Sct should be undefined, since νt = 0 and Dt = 0 in laminar
Throughout the last century, the turbulent Schmidt (Sct) and regions. The boundedness of eq 20 in laminar regions makes its use
Prandtl (Prt) numbers have played a key role in turbulent passive practical, from a modeling perspective, albeit incorrectly overstating
scalar transport modeling that utilizes the gradient diffusion turbulent mass flux in the laminar sublayer of boundary layers. Finally,
hypothesis (eq 15). Specifically, Sct or Prt are used to relate the all of the relationships presented by Reynolds classified the Sct and Prt
turbulent mass or thermal diffusivity in eq 15 to the turbulent numbers into Prandtl mixing-length-based models, simple empiri-
viscosity (νt) determined by a turbulent-viscosity model (eq 5) in cism, and statistical calculations. Reynolds concluded that the mixing
the RANS or URANS equations. This leads to the key question length models were not fundamentally advantageous, since they do
of what determines Sct and Prt and why are they important? This not elucidate transport phenomena but are practical in the sense that
is briefly addressed in the following section. they are easy to use and contain realistic information. The remaining
In general, Sct or Prt is treated as a global parameter in complex models were neither practical nor fundamentally advantageous, since
fluid simulations; the parameter is usually set to a default value of they merely served to “remind us how little confidence can be placed
0.7 or unity. The treatment of Sct and Prt numbers in this sense is in any limited group of measurements”.29 Although the mixing-length
ubiquitous: they are used in both open source18 and commercial19 models were considered most desirable at the time, Reynolds
CFD packages; are accepted in scientific literature;3,2022 and concluded that more work should be done to account for position
detailed in the scientific literature by Pope,5 Ranade,23 and Fox11 from the wall separately from the turbulent intensity.
devoted to the subject of reacting turbulent flows and chemical Since the review of A. J. Reynolds,29 further research related to
reaction engineering. Estimates of a global turbulent Schmidt the prediction of Sct and Prt numbers has been more empirical.
number in homogeneous turbulent flows were discussed thor- Work by Jischa30 described the parameters Prt and Sct, using
oughly by Corrsin24 by comparing Taylor microscales for the simple empirical correlations,
scalar φu and velocity uu components. As noted by Corrsin, values B
of Sct were approximately unity or less, in agreement with the work Pr t ¼ A þ
of Batchelor.25,26 Although the results of Corrsin were valuable for Pr
homogeneous turbulence, inhomogeneous turbulence is often B
encountered in practice and requires transport equations to model Sct ¼ A þ ð21Þ
its effects. Nevertheless, the use of a constant global turbulent Sc
Schmidt number in packed-bed simulations,22 urban diffusion with no influence of wall distance or turbulent intensity taken into
problems,3 and combustion modeling27,28 is seen throughout account. A review by Kays31 examined experimental data on the Prt
computational fluid dynamics, ranging in value from 0.1 to 2.2. number for 2D turbulent boundary layer flows in circular pipes or
This large range alludes to the fact that prescribing a global value is flat ducts. Kays discussed empirical models based on DNS and
8820 dx.doi.org/10.1021/ie200055s |Ind. Eng. Chem. Res. 2011, 50, 8817–8823
Industrial & Engineering Chemistry Research REVIEW

experimental data to predict Prt in different regions of the boundary This is true with respect to the Prt number in a published work by
layer for various fluids (i.e., air, water, oil, liquid metals). The Churchill in 2002.16 Churchill provided a reinterpretation of the
conclusions by Kays further support the conclusions by A. J. Prt number, for a fully developed turbulent flow in a round tube,
Reynolds29 in that the Prt and Sct numbers are dependent on the in terms of a local fraction of shear stress and fraction of heat flux
molecular Pr and Sc numbers and the distance from the wall. More density due to fluctuations in velocity. The essential steps in the
recent work by Koeltzsch32 investigated the height dependence of derivation require the stresses in the radial direction, the trans-
Sct in boundary layers through wind-tunnel experiments. As a result, port of energy in the negative radial direction, and a formula-
a power series approximation of Sct was used to fit experimental data, tion of dimensionless shear and heat flux density. As a result,
only to further show the wall dependence of Sct in turbulent Churchill was able to formulate
boundary layer flows. In a different approach, Guo et al.33 applied  
the genetic algorithm to optimize a variable Sct proportional to three Pr t ðu0 v0 Þþþ 1  ðT 0 v0 Þþþ
constants to control the magnitude of the Sct, the relative impor- ¼   ð23Þ
tance of turbulent frequency scale, and the affect of asymmetry in the
Pr ðT 0 v0 Þþþ 1  ðu0 v0 Þþþ
stresses to agree with experimental results for a jet in crossflow.
Validated against three different cases, Guo et al. showed qualitative Equation 23 expands the Prandtl numbers in terms of dimen-
and quantitative agreement with previously published data. For sionless shear stresses,
completeness, a comparison against cases of constant global Sct was 0 0
performed. The work by Guo et al. addresses the need for new FÆui uj æ
ðu0 v0 Þþþ ¼ ð24Þ
spatially and temporally variable Sct (rather than a global constant τij
Sct) and shows that a variable Sct number fits numerical data to
experimental data better than using a constant Sct number, but fails and dimensionless heat flux density,
to elucidate phenomena influencing the Sct in turbulent flows.
FcÆT 0 u0 æ
The overall importance of the Sct and Prt parameters in CFD is ðT 0 v0 Þþþ ¼ ð25Þ
that the value of these two numbers strongly influence the level of q
turbulent mixing in the modeled system. If small global Sct and
where τij is the total shear in the radial direction, q the heat flux
Prt numbers are assumed, then a large turbulent mass or thermal
density in the y-direction, and F the fluid density, and c the heat
diffusivity is defined. For instance, in combustion, a lower Sct
capacity.
number intensifies combustion due to enhanced species diffusion
Equation 23 is an important starting point for relating Prandtl
and turbulent mixing, while a higher Sct number may create
numbers in fully developed turbulent flow in channels and pipes.
mixtures that are unable to sustain combustion.27 Without a
One can see the possibilities of extending this methodology to
more fundamental understanding of the Sct and Prt numbers,
spatially variable Prt (or Sct) in more-complex geometries. The
more uncertainty is introduced into CFD simulations.
crux of using such an approach is the determination of (uuvu)þþ
The implications of an improved ability to predict Prt and Sct
and (Tuvu)þþ. Based on the work by Papavassiliou and Hanratty,36
reach beyond CFD to more classical fields of mass- and heat-
Churchill noted that a Lagrangian form of DNS provided an
transfer correlations using the Sherwood and Nusselt numbers
accurate calculation of these quantities without empiricism. Also, a
(Sh and Nu, respectively). For the turbulent Prt number, a review
recent work by Srinivasan and Papavassiliou37 explored the use of
by Churchill in 200034 noted that
Lagrangian DNS to determine Prt for classical Poiseulle channel
“The development of a comprehensive predictive or corre- and plane Couette flows. Prt was found to be a function of Pr,
lative expression for the turbulent Prandtl number is the which agrees with the observations presented earlier in this paper.
principle remaining challenge with respect to the prediction The extension of determination of Prt for curved geometries
of turbulent forced convection.” (where transport is not necessarily unidirectional) remains an
open challenge and is the subject of ongoing research in our
This can be equally stated for the Sct number in highly complex, laboratory that will be addressed in upcoming publications.
boundary layer, and multiscale flows. Dudukovic and Pjanovic35
found that “it is precisely this assumption of Sct = (constant) that
leads to misrepresentation of the Reynolds number depen- 8. CONCLUSIONS
dence”, referring to a Sherwood number correlation, As a precursor to passive scalar transport modeling, the
Reynolds-averaged NavierStokes (RANS) equations, along with
rffiffiffiffiffiffi
f Sc the turbulent closure issue, were provided as a starting point.
Sh ¼ a1 Re ð22Þ Inasmuch the RANS equation requires closure to account for the
Sct
effect of chaotic turbulent fluctuations, so does the transport of
a passive scalar in a turbulent flow field. Using Reynolds averaging,
for boundary layers in pipe flows and falling films. The conclu-
sions by Dudukovic and Pjanovic further stated that Sct is not a scalar is broken into ensemble-averaged and fluctuating compo-
nents (eq 11), while creating an additional unclosed scalar-flux
constant and is dependent on turbulent spectra, which agrees
term ÆuuφuRæ. We presented several methods used to close the
with the conclusions by Reynolds29 concerning the importance
scalar-flux term, including the gradient diffusion hypothesis
of the turbulent intensity in predicting the Sct value.
(GDH) (eq 15), algebraic models (eqs 17 and 18), and a scalar-
flux transport model (eq 19).
7. CHURCHILL’S REINTERPRETATION AND A LOOK TO Of the models presented, the GDH is the simplest and least
THE FUTURE computationally intensive. The only requirement is the determi-
Re-examination from a different perspective often leads to nation of a turbulent mass or thermal diffusivity, via a turbulent
improvements in fundamental knowledge and understanding. Schmidt Sct or Prandtl number Prt and a known turbulent
8821 dx.doi.org/10.1021/ie200055s |Ind. Eng. Chem. Res. 2011, 50, 8817–8823
Industrial & Engineering Chemistry Research REVIEW

viscosity (νt) calculated by an eddy viscosity turbulence model. It a1 = model constant used in eq 22 (from ref 35)
is common practice to define a constant global Sct or Prt as a flow- B = model constant used in eq 21 (from ref 30)
specific (e.g., combustion, urban diffusion, jet in crossflow) value C1 = model constant used in eq 20 (from ref 29)
usually in the range of 0.1 to >1 for both dimensionless numbers. C2 = model constant used in eq 20 (from ref 29)
However, by setting a constant global Sct or Prt, additional Cθ = DalyHarlow model constant used in eq 17 (from ref 13)
uncertainty is introduced into the numerical CFD simulation c = heat capacity at constant pressure
through assumptions about levels of turbulent mass or heat D = mass diffusivity
diffusion and turbulent mixing in the system. For example, Dtij = anisotropic turbulent mass diffusivity
combustion simulations have shown that an assignment of lower f = model constant used in eq 21 (from ref 35)
Sct intensifies combustion, because of enhanced species diffusion J = molecular flux (eq 19)
and turbulent mixing, while a higher Sct number may create k = turbulent kinetic energy
mixtures that are unable to sustain combustion. Pr = Prandtl number; Pr = ν/R
What has been shown in previous literature is that both the Sct P = production term
and Prt numbers are strongly influenced by three factors: the p = pressure
corresponding molecular value of the Pr and Sc parameters, the q = heat flux density
position within the flow (i.e., distance from the wall), and the R = pressure-scalar-gradient term
local turbulent intensity of the flow. This indicates that spatially S = scalar source term (eq 14)
and temporally variable Prt and Sct are more appropriate and Sc = Schmidt number; Sc = ν/D
accurate in approximating the turbulent diffusivity in eq 15. Sh = Sherwood number (eq 22)
Furthermore, research has shown that a spatially variable Sct model t = time
produced simulation results in much closer agreement with U = velocity vector
experimental data than simulations using a global Sct number. u = fluctuating velocity component
Until recently, models predicting the Sct or Prt number have been (uuvu)þþ = dimensionless shear stresses;
limited to mixing-length models, empirical relations, statistical (uuvu)þþ = FÆuuvuæ/τ (eq 24)
þþ
calculations, or optimization algorithms. Work by Professor Stuart (Tuvu) = dimensionless heat flux density;
Churchill has provided further insight that supports the knowledge (Tuvu)þþ = FcÆTuuuæ/q (eq 25)
that fluctuations in velocity—more importantly, the amount of x = Cartesian vector space
dimensionless shear—play an important role in heat transport and x = independent space variable
may be used to relate molecular and turbulent Prandtl numbers Æ 3 æ = ensemble average
within fully developed turbulent flows in a pipe. What can be
reiterated is that an improved knowledge of Sct and Prt is necessary Greek Symbols
to fully exploit the GDH, especially in situations where more- r = thermal conductivity
intricate modeling techniques (i.e., large eddy simulation (LES), R = species
direct numerical simulation (DNS), scalar-flux models) are in- δij = Kronecker delta
tractable, because of computational domain complexities and/or ε = turbulent or scalar-flux dissipation
lack of computational resources. μ = viscosity
In summary, the development of a comprehensive approach ν = kinematic viscosity
for evaluation of the turbulent Prandtl (or Schmidt) number Π = velocitypressure gradient
remains as much of a principal challenge, with respect to the F = density
prediction of scalar transport in turbulent forced convection, τ = total shear stress tensor
today as it was a decade ago, as indicated by Churchill. The φ = a conserved scalar
increasing demands for improved efficiency in the energy and ω = specific turbulent dissipation rate
chemical and materials processing areas provide the incentives
Subscripts
for renewed efforts.
i, j, k = roman indices notation
t = turbulent
’ AUTHOR INFORMATION T = total
Corresponding Author
*E-mail: dudu@wustl.edu. Accents/Superscripts
u = fluctuations
m, n = model exponents used in eq 20 (from ref 29)
’ ACKNOWLEDGMENT t = anisotropic turbulence
We are grateful for the work by Professor Churchill and his unique Acronyms
insight into turbulence modeling. His work has been useful to us in CFD = computational fluid dynamics
our research in multiphase modeling as well as the teaching of DNS = direct numerical simulation
transport phenomena to the future generation of engineers. We GDH = gradient diffusion hypothesis (eq 15)
would also like to thank the industrial partners of the Chemical LES = large eddy simulation
Reaction Engineering Laboratory (CREL) at Washington University RANS = Reynolds-averaged NavierStokes equations
for their generous support of our research projects. RNG kε = renormalization group kε turbulence model
SGS = LES subgrid scale models
’ NOMENCLATURE URANS = unsteady Reynolds-averaged NavierStokes equation
A = model constant used in eq 21 (from ref 30) (see eq 4)
8822 dx.doi.org/10.1021/ie200055s |Ind. Eng. Chem. Res. 2011, 50, 8817–8823
Industrial & Engineering Chemistry Research REVIEW

URAPS = Reynolds-averaged passive scalar equation (see eq 14) (28) Liu, Y.; Feng, H.; Olsen, M. G.; Fox, R. O.; Hill, J. C. Turbulent
kω = shear stress transport (SST) k-omega model mixing in a confined rectangular wake. Chem. Eng. Sci. 2006,
61, 6946–6962.
(29) Reynolds, A. The prediction of turbulent Prandtl and Schmidt
’ REFERENCES numbers. Int. J. Heat Mass Transfer 1975, 18, 1055–1069.
(1) Dimotakis, P. E. Turbulent Mixing. Annu. Rev. Fluid Mech. 2005, (30) Jischa, M.; Rieke, H. B. About the prediction of turbulent
37, 329–356. prandtl and schmidt numbers from modeled transport equations. Int. J.
(2) Warhaft, Z. Passive Scalars in Turbulent Flows. Annu. Rev. Fluid Heat Mass Transfer 1979, 22, 1547–1555.
Mech. 2000, 32, 203–240. (31) Kays, W. M. Turbulent Prandtl number. Where are we? ASME
(3) Tominaga, Y.; Stathopoulos, T. Turbulent Schmidt numbers for Trans. J. Heat Transfer 1994, 116, 284–295.
CFD analysis with various types of flowfield. Atmos. Environ. 2007, (32) Koeltzsch, K. The height dependence of the turbulent Schmidt
41, 8091–8099. number within the boundary layer. Atmos. Environ. 2000, 34, 1147–
(4) Bird, R. B.; Stewart, W. E.; Lightfoot, E. N. Transport Phenomena, 1151.
2nd ed.; Wiley: New York, 2001. (33) Guo, Y.; He, G.; Hsu, A. T. Application of genetic algorithms to
(5) Pope, S. B. Turbulent Flows, 1st ed.; Cambridge University Press: the development of a variable Schmidt number model for jet-in-cross-
Oxford, U.K., 2000. flows. Int. J. Numer. Methods Heat Fluid Flow 2001, 11, 744–761.
(6) Aris, R. Vectors, Tensors and the Basic Equations of Fluid (34) Churchill, S. W. Progress in the thermal sciences: AIChE
Mechanics; Dover Publications: New York, 1989. Institute Lecture. AIChE J. 2000, 46, 1704–1722.
(7) Boussinesq, J. Theorie de l’ecoulement tourbillant. Mem. Pres. (35) Dudukovic, A.; Pjanovic, R. Effect of Turbulent Schmidt
Divers Savants Acad. Sci. Inst. Fr. 1877, 23, 46–50. Number on Mass-Transfer Rates to Falling Liquid Films. Ind. Eng.
(8) Pletcher, R. H.; Tannehill, J. C.; Anderson, D. A. Computational Chem. Res. 1999, 38, 2503–2504.
Fluid Mechanics and Heat Transfer, 2nd ed.; Taylor & Francis: (36) Papavassiliou, D. V.; Hanratty, T. J. Transport of a passive scalar
Washington, DC, 1997. in a turbulent channel flow. Int. J. Heat Mass Transfer 1997, 40, 1303–
(9) Ferziger, J. H.; Peric, M. Computational Methods for Fluid 1311.
Dynamics, 3rd ed.; Springer: Berlin, New York, 2002. (37) Srinivasan, C.; Papavassiliou, D. V. Prediction of the Turbulent
(10) Wilcox, D. C. Turbulence Modeling for CFD, 3rd ed.; DCW Prandtl Number in Wall Flows with Lagrangian Simulations. Ind. Eng.
Industries: La C~anada, CA, 2006. Chem. Res. 2010, DOI: 10.1021/ie1019497.
(11) Fox, R. O. Computational Models for Turbulent Reacting Flows;
Cambridge University Press: Oxford, U.K., 2003.
(12) Batchelor, G. K. Diffusion in a Field of Homogeneous Turbu-
lence. I. Eulerian Analysis. Aust. J. Sci. Res. A 1949, 2, 437.
(13) Younis, B. A.; Speziale, C. G.; Clark, T. T. A rational model for
the turbulent scalar fluxes. Proc. R. Soc., Ser. A 2005, 461, 575.
(14) Daly, B. J.; Harlow, F. H. Transport Equations in Turbulence.
Phys. Fluids 1970, 13, 2634–2649.
(15) Kim, J.; Moin, P.; Moser, R. Turbulence statistics in fully
developed channel flow at low Reynolds number. J. Fluid Mech. 1987,
177, 133–166.
(16) Churchill, S. W. A Reinterpretation of the Turbulent Prandtl
Number. Ind. Eng. Chem. Res. 2002, 41, 6393–6401.
(17) Levich, V. G. Physicochemical Hydrodynamics; Prentice Hall:
Englewood Cliffs, NJ, 1962.
(18) Weller, H. G.; Tabor, G.; Jasak, H.; Fureby, C. A tensorial
approach to computational continuum mechanics using object-oriented
techniques. Comput. Phys. 1998, 12, 620–631.
(19) Ansys Fluent 6.3 User’s Guide; Ansys, Inc.: Houston, TX, 2006.
(20) Jiang, L.; Campbell, I. Prandtl/Schmidt number effect on
temperature distribution in a generic combustor. Int. J. Therm. Sci.
2009, 48, 322–330.
(21) Lipatnikov, A.; Chomiak, J. Effects of premixed flames on
turbulence and turbulent scalar transport. Prog. Energy Combust. Sci.
2010, 36, 1–102.
(22) Frigerio, S.; Thunman, H.; Leckner, B.; Hermansson, S. Esti-
mation of gas phase mixing in packed beds. Combust. Flame 2008,
153, 137–148.
(23) Ranade, V. Computational Flow Modeling for Chemical Reactor
Engineering, 1st ed.; Academic Press: New York, 2001.
(24) Corrsin, S. The isotropic turbulent mixer: Part II. Arbitrary
Schmidt number. AIChE J. 1964, 10, 870–877.
(25) Batchelor, G. K. Small-Scale Variation of Convected Quantities
Like Temperature in Turbulent Fluid, Part 1. General Discussion and
the Case of Small Conductivity. J. Fluid Mech. 1959, 5, 113–133.
(26) Batchelor, G. K.; Howells, I. D.; Townsend, A. A. Small-Scale
Variation of Convected Quantities Like Temperature in Turbulent
Fluid, Part 2. The Case of Large Conductivity. J. Fluid Mech. 1959,
5, 134–139.
(27) Baurle, R. A. Modeling of high speed reacting flows: established
practices and future challenges. AIAA J. 2004, 267, 42.

8823 dx.doi.org/10.1021/ie200055s |Ind. Eng. Chem. Res. 2011, 50, 8817–8823

Вам также может понравиться