You are on page 1of 20

European Conference on Computational Fluid Dynamics ECCOMAS CFD 2006 P. Wesseling, E. O nate and J.

P eriaux (Eds) c TU Delft, The Netherlands, 2006

CFD MODELING OF CHEMICAL REACTORS: SINGLE-PHASE COMPLEX REACTIONS AND FINE-PARTICLE PRODUCTION
Ying Liu, Qing Tang and Rodney O. Fox
Department

of Chemical and Biological Engineering, Iowa State University Ames, IA 50010-2230, U.S.A. e-mail: liuying@iastate.edu, rofox@iastate.edu Reaction Engineering International Salt Lake City, UT 84101, U.S.A. e-mail: tang@reaction-eng.com

Key words: Computational Fluid Dynamics, Chemical Reactors, Turbulent Reacting Flow, Quadrature Method of Moments Abstract. Computational uid dynamics (CFD) is a useful tool for modeling chemical reactors. However, because the design goals and expected outcomes are dierent than in traditional CFD applications, chemical reactors require special attention to the treatment of chemical reactions, and heat and mass transfer. Here we provide an overview of the modeling components needed to describe single-phase reactors with complex reactions and possible ne-particle production, and show some successful examples from our laboratory. The models are described in the context of the Reynolds-average transport equations, but can be easily modied for use with large-eddy simulations. The examples range in complexity from turbulent mixing of a single scalar to turbulent reacting ow with the formation of ne particles. For the latter, we illustrate how the number density function describing the particle population can be eciently integrated with a CFD code by using the quadrature method of moments.

INTRODUCTION

Computational uid dynamics (CFD) is most often associated with solving for the uid dynamics in single and multiphase ows. While this is certainly a critical part of modeling chemical reactors, it is only the starting point. When developing CFD models for chemical reactors much of the eort is directed at correctly describing the chemical and physical processes that inuence the productivity of the reactor. For example, chemical reactions and heat and mass transfer must be accurately modeled in order to predict the selectivity and yield of mixing-sensitive reactions. Depending on the reactor type (single vs. multiphase, gas vs. liquid, etc.), the rate-controlling steps that must be included in 1

Ying Liu, Qing Tang, and Rodney O. Fox

the CFD model can be very dierent. Thus, unlike in more traditional applications of CFD, this makes CFD modeling of chemical reactors more challenging (and perhaps more interesting) for the engineers responsible for developing the model. In this work we discuss CFD modeling of single-phase chemical reactors under turbulent ow conditions, which includes reactors for making ne particles. For such reactors, there are several key components in the CFD model that inuence the accuracy of the predictions: The turbulence model that is used to describe the uid dynamics. The scalar transport model that is used to describe turbulent mixing. The model used to describe the eect of turbulence on chemical reactions, and vice versa. The transport model used to describe the evolution of ne particles due to nucleation, growth, aggregation, etc. As we will see in the following discussion, each of these modeling components can be added sequentially to arrive at the overall CFD model. Likewise, each of them can (and should) be validated experimentally with appropriate measurements (i.e. the turbulence statistics should be measured directly and not inferred indirectly from, for example, measurements of product yield.) Likewise, because the chemical kinetics and particle evolution processes will have a strong inuence on the CFD predictions, it will be important that models for these processes be adequately tested in simplied ow congurations before they are used in CFD models of plant-scale chemical reactors. The remainder of this work is arranged as follows. First we discuss CFD models for turbulent mixing and show how such models can be validated with experiments. Next we add a micromixing model for mixing-sensitive reactions and apply it to a liquid-phase conned impinging-jets reactor to illustrate how CFD can be used to understand mixing in such devices. In the third example, we consider turbulent combustion wherein the chemical reactions aect the ow eld through the density eld. Finally, in the last example we show how the number density function needed to describe a population of ne particles can be included in CFD models for chemical reactors. All of the examples use models developed at Iowa State, but we should point out that there are many other examples in the literature that could also be cited. Our goal here is not to give an exhaustive overview, but rather to illustrate the necessary steps and concepts required to develop an accurate CFD model. 2 TURBULENT MIXING

Most chemical reactors are designed to operate in the turbulent regime to maximize throughput. It follows that a detailed understanding of turbulent mixing is crucial for proper design and optimization of chemical reactors, inspiring numerous experimental 2

Ying Liu, Qing Tang, and Rodney O. Fox

and computational studies on turbulent mixing over the years 1,2,3. In addition to describing large-scale segregation (sometimes called macromixing or blending), a successful CFD mixing model should be able to accurately account for the complex interactions between turbulence and mixing at the sub-grid scale 4 . The latter is often referred to as micromixing or molecular-scale mixing to emphasize its intimate connection with molecular diusion and to distinguish it from large-scale mixing processes that are convective in nature. However, when reading the chemical engineering literature on mixing one must be careful when interpreting the meaning of the term micromixing as it is now used to describe mixing in microreactors. Even in the older literature on reactor modeling the term micromixing is often used to describe all mixing processes that are not captured at the level of modeling being used. For example, a stirred tank might be modeled as well macromixed and any concentration uctuations about the tank-average mean concentration are treated using a micromixing model. In the context of CFD, it is preferable to employ a tighter denition of micromixing based on the concentration uctuations about the pointwise local average concentration 4 (in other words, uctuations that are not resolved by the transport equation for the Reynolds-averaged or ltered scalar eld.) Unlike for gas-phase mixing (e.g. turbulent jets and scalar mixing layers), there exists very little data in the literature on liquid-phase mixing that can be used for CFD model validation. Thus at Iowa State we have built a ow facility with detailed laser-based diagnostics to develop such databases. For example, we have recently reported on CFD validation eorts using the turbulent ow eld and the inert scalar (i.e. mixture fraction) eld in a conned planar-jet reactor. (The ow facility and experimental techniques are described in detail elsewhere 5.) The conned planar-jet reactor has a cross section of 0.06 0.1 m2 and an overall length of 1 m. Three rectangular inlet jets of width 0.2 m enter the reactor at ow rates 0.5, 1, 0.5 m/s, respectively. The turbulence and scalar elds are measured by two non-intrusive optically based techniques: particle-image velocimetry (PIV) and planar laser-induced uorescence (PLIF). Our in-house CFD code is a hybrid nite-volume (FV) Reynolds-average Navier-Stokes (RANS)/transported probability density function (PDF) code 5 . The FV code supplies the transported PDF code with ow statistics such as the mean velocity, turbulent kinetic energy. The Reynolds stresses are closed by a two-layer k - model 6 . The FV code also solves the Reynolds-averaged scalar moment (rst and second) transport equations with the ux terms and the scalar dissipation closed by the gradient-diusion model and the equilibrium model, respectively. The PDF code solves the scalar eld represented by Lagrangian particles. For consistency, the scalar-ux term in the PDF model is approximated by the gradient-diusion model, and the micromixing term is represented by the interaction-by-exchange-with-the-mean (IEM) model. The experimental data for the conned jet were used to validate the turbulent mixing models implemented in our CFD code. The proles of the mixture-fraction mean predicted by the RANS model and the transported PDF method with dierent values of the 3

Ying Liu, Qing Tang, and Rodney O. Fox


1

0.8

0.8

0.8

0.6

0.6

0.6

<>

<>

<>
0.4

0.4

0.4

0.2

0.2

0.2

0 -1.5

-0.75

0.75

1.5

0 -1.5

-0.75

0.75

1.5

0 -1.5

-0.75

0.75

1.5

y/d

y/d

y/d

Figure 1: Comparison of mean mixture fraction proles for (a) x/d = 1, (b) x/d = 7.5, (c) x/d = 15. , PLIF; , RANS, ScT = 0.7; - - -, RANS, ScT = 0.5; , PDF, ScT = 0.7; , PDF, ScT = 0.5.
0.1

0.1

0.1

0.08

0.08

0.08

0.06

0.06

0.06

< >

< >

< >
0.04 0.02 0 -1.5

,2

,2

0.04

0.04

0.02

0.02

0 -1.5

-0.75

0.75

1.5

0 -1.5

,2

-0.75

0.75

1.5

-0.75

0.75

1.5

y/d

y/d

y/d

Figure 2: Comparison of mixture-fraction variance proles (a) x/d = 1, (b) x/d = 7.5, (c) x/d = 15. , PLIF; , RANS, ScT = 0.7; - - -, RANS, ScT = 0.5; , PDF, ScT = 0.7; , PDF, ScT = 0.5.

turbulent Schmidt number ScT are presented in Fig. 1, and compared with those measured by PLIF at three downstream locations. They agree quite well, indicating that the turbulence model and the gradient-diusion model accurately predict the scalar transport for this ow geometry. The lower spreading rate of the mixture-fraction mean in the simulations suggests that the turbulent Schmidt number is slightly lower than the typical value 0.7. The agreement improves when ScT is reduced to 0.5. We should note that in order to achieve good agreement for the mean mixture fraction proles, it is very important that the mean velocity and turbulence elds be correctly predicted by the CFD model. As shown elsewhere 5, the two-layer k - model yields excellent predictions for this ow conguration. Similarly, Fig. 2 shows the comparison of the mixture-fraction variance elds predicted by the RANS equation as well as the transported PDF method and the experimental data at various streamwise locations. The RANS equation and the PDF method yield quite similar results except at x/d = 1 where a relatively small time step is required for the PDF method due to the slow turbulent mixing. In general, the CFD predictions in the shear layers are higher than the PLIF data, and even a smaller ScT can not improve the agreement. The equilibrium mixing model (which neglects the transport terms for

Ying Liu, Qing Tang, and Rodney O. Fox

the scalar dissipation rate) and the limited spatial resolution of the PLIF measurements jointly account for the observed discrepancy. Through the theory of the scalar energy spectrum 4 , the mixture-fraction variance missed by PLIF can be estimated quantitatively (Table 1) 5 . For this ow, the Batchelor length scale grows along the streamwise direction, meaning PLIF resolves more sub-grid scalar-energy-containing eddies even though the spatial resolution remains constant. Therefore the variance not resolved by PLIF decreases at further downstream locations. This tendency is consistent with that shown by Fig. 2. x/d 1 variance unresolved by PLIF (%) 12.68 4.5 9.36 7.5 8.01 12 7.75 15 5.98

Table 1: Estimated variance unresolved by PLIF at the peak of the variance prole.

Overall, our experience with CFD models for liquid-phase turbulent mixing has shown that, provided the ow is suciently turbulent, the existing CFD models based on Reynolds-average quantities work well. On the other hand, we have also tested largeeddy simulations (LES) for the conned planar-jet conguration with much less success. For example, the LES results for the mean velocity eld (regardless of the sub-grid scale model used for the stresses) are signicantly dierent than the PIV experiments (and the two-equation RANS model). We attribute this dierence to the inadequacy of the treatment of wall boundary layers in LES, which play a strong role as the planar jet transitions to channel ow in our experiments. Unfortunately, the performance of the LES wall model 7 that we have tested gets worse at higher Reynolds numbers unless the wall boundary layers are signicantly resolved (which makes the LES quite expensive). In our opinion, these questions merit further research and one should be very careful and not assume that standard LES models will necessarily yield better (or even as good as) results as RANS models for conned ows. 3 MIXING-SENSITIVE REACTIONS

One of the important uses of CFD in chemical reactor modeling is to predict the outcomes of chemical reactions and their dependence on turbulent mixing. Transported PDF methods are well known for their ability to treat the interactions between turbulence and chemistry in detail with good to high accuracy 8,9. However, the relatively high computational cost of methods in this category has motivated us to explore lower-cost methods for approximating the transport equation for the joint scalar PDF 10 . Fox 4 proposed the direct quadrature method of moments (DQMOM), which approximates an Eulerian joint PDF given its closed transport equation. By applying DQMOM to the composition PDF transport equation closed by the IEM model and assuming the form of the PDF as a multi-peak delta function, the model equations of the multi-environment

Ying Liu, Qing Tang, and Rodney O. Fox

DQMOM-IEM model are obtained: Dpn = (T pn ) , (1) Dt Dpn n = [T (pn n )] + pn ( n ) + pn S (n ) + bn , (2) Dt where is the uid density, pn is the mass fraction of the nth (n = 1, , Ne ) environment. n is the composition (e.g. molar concentration or mass fraction) of the chemical species ( = 1, 2, , Ns ) in the nth environment and is its mean composition. The convected derivative D/Dt involves the mean velocity, which can be computed using LES or RANS methods. Likewise, the turbulent diusivity T is found by solving an appropriate turbulence model. As with all PDF methods, in Eq. 2 the chemical source term S () appears in closed form. The rate of micromixing between environments is controlled by . The correction term bn arises due to diusion in real space in the presence of a mean scalar gradient and can be computed by solving the following system of linear equations for each chemical species ( = 1, . . . , Ns ):
Ne 1 m n bn = n=1 n=1 Ne 2 (m 1) pn m n T |n | 2

for m = 1, . . . , Ne .

(3)

For non-premixed ows, the boundary conditions for pn determine which uid is present. For example, with two inlets (Ne = 2) we will have p1 = 1 in the rst inlet and p2 = 1 in the second inlet, while p1 + p2 = 1 at every point in the ow. Thus, physically, p1 (x) is the mass fraction of uid at location x that entered the reactor through the rst inlet. The boundary conditions for n then determine the chemical compositions in each inlet stream. When the uid is completely mixed on all scales, p1 equals the mass fraction of uid entering the reactor through the rst inlet stream. In comparison to transported PDF methods, it is important to note that the DQMOM-IEM model has no statistical error (or bias). Thus, for example, even with just one environment the scalar mean (in the absence of chemistry) can be predicted accurately. On the other hand, transported PDF methods use more notional particles (partially to reduce statistical errors), and thus should provide a more accurate representation of highly nonlinear chemical source terms. The exact trade-o between the two methods is thus dependent on the application. More details on multi-environment CFD models can be found elsewhere 4. The DQMOM-IEM model has been employed to model the conned impinging-jet reactor (CIJR) from the experimental study reported by Johnson and Prudhomme 11. Non-premixed reactants were introduced into the CIJR with two co-axial impinging jets at equal ow rates. The mixing times produced are on the order of milliseconds and decrease with the increase of the inlet Reynolds number Rej . A mixing-sensitive, consecutivecompetitive reaction system was used to investigate the reactive mixing 11:
1 H+ + OH H2 O 2 H+ + CH3 C(OCH3)2 CH3 (DMP) + H2O H+ + CH3 COCH3 + 2CH3 OH

(4)

Ying Liu, Qing Tang, and Rodney O. Fox

Following the procedures outlined in Fox 4, the concentrations of the chemical species in Eq. 4 can be written in terms of the mixture fraction and two reaction-progress variables Y1 and Y2 . Since k1 k2 , the rst reaction is controlled by mixing and thus Y1 is determined exclusively by . Consequently, only and Y2 are of interest. Moreover, in all the experiments and computations, the reactant OH is in excess, enabling the conversion of DMP to be a sensitive measure of the quality of mixing in the reactor. Consistently, the reaction time scale tr is represented by the slow reaction and the inlet concentration of H+ , which catalyzes the slow reaction. For a two-environment DQMOM-IEM model, the variables to be modeled are p1 , 1 , 2 , Y21 and Y22 by assuming that one inlet stream contains only the rst environment in which the mixture fraction (1 ) is zero and the other inlet stream contains only the second environment in which the mixture fraction (2 ) equals 1. Since the reactants are non-premixed, no reaction happens in the inlet streams (Y21 = Y22 = 0). The Reynolds-average concentration of each species is the probability average of the concentration of that species in each environment. For example, the mean 2 2 mixture fraction is = p1 1 + p2 2 , and the second moment is 2 = p1 1 + p2 2 . We should note that solving transport equations for p1 , 1 , and 2 has several numerical advantages (e.g., the variance and higher-order moments will always be realizable) over solving the moment transport equations directly. Readers interested in details on the reactor geometry and the CFD model can refer to Liu & Fox 12. For this reactor, the turbulent ow is not fully developed for all of the jet Reynolds numbers considered in the experiments. To compute the turbulence elds, we again use a two-layer k - model. For this particular ow, we cannot validate the velocity and turbulence elds directly as no experimental data are available. Instead, the CFD model predictions for the conversion of DMP in the outlet stream were compared to experiments. Due to the relatively low Reynolds number, the eects of the Reynolds number (and Schmidt number) on the micromixing rate (i.e. on ) must be taken into consideration in order to get accurate CFD predictions. For this purpose, the dependence of the mechanical-to-scalar time-scale ratio on the local turbulent Reynolds number was approximated by a polynomial 12. It turns out that accounting for the Reynolds-number dependence of was the key element in the success of the CFD model at predicting the experimental data. Figure 3 shows the distribution of the Reynolds-average concentration of H+ , OH and DMP in a cross-section of the reactor, which includes the inlet and outlet tubes for Rej = 400 and tr = 81 ms. H+ enters the CIJR in the left inlet stream while OH and DMP are contained in the right inlet stream. H+ is consumed completely and residual amounts of OH exit in the outow. The mixing-cup average concentration of DMP at the outlet is below the concentration after complete mixing by a few percent due to the reaction catalyzed by H+ . In fact, the most intensive conversion of DMP occurs near the left wall of the CIJR where H+ is in slight excess. Since the macromixing beyond the impingement area is relatively weak, the DMP is not uniformly distributed across the outlet tube, neither are the mixture fractions nor the reaction-progress variables in the 7

Ying Liu, Qing Tang, and Rodney O. Fox

5 4 3 2
Z(mm) Z(mm)

5 4 3 2 1 0 -1 -2 2 -2 -1 0
X(mm)
OH (M)
-

5 4 3 2
Z(mm)

1 0
H (M)
49 44 40 35 31 27 22 18 13 9 4 0
+

1 0
DMP (M)
49 44 40 35 31 27 22 18 13 9 4 0

-1 -2 -2 -1 0
X(mm)

51 46 42 37 32 28 23 19 14 9 5 0

-1 -2 2 -2 -1 0
X(mm)

Figure 3: Reynolds-average species distributions for Rej = 400 and tr = 61 ms.

two environments (Fig. 4). Thus, the reactions will continue along the outlet tube. This problem can be avoid by adjusting the ratio of the inlet concentration of H+ to OH . More analysis and optimization suggestions can be found in Liu & Fox 12.
0.5
1

0.5

2
0.530 0.525 0.519 0.514 0.508 0.503 0.497 0.492 0.486 0.481 0.475 0.470

0.5

Y21

0.5

Y22
0.600 0.569 0.538 0.507 0.476 0.445 0.415 0.384 0.353 0.322 0.291 0.260

0.5

DMP (M)
240 227 214 200 187 174 161 148 135 121 108 95

Y(mm)

Y(mm)

Y(mm)

Y(mm)

Y(mm)

-0.5 -0.5 0.5

0 X(mm)

0.5
1

-0.5 -0.5 0.5

0 X(mm)

0.5
2

-0.5 -0.5 0.5

0 X(mm)

0.5
Y21

-0.5 -0.5 0.5

0 X(mm)

0.5
Y22

-0.5 -0.5 0.5

0 X(mm)

0.5
DMP (M)

Y(mm)

Y(mm)

Y(mm)

Y(mm)

Y(mm)

-0.5

0 X(mm)

0.5

-0.5

0 X(mm)

0.5

-0.5

0 X(mm)

0.5

-0.5

0 X(mm)

0.5

-0.5 -0.5

0 X(mm)

0.5

Figure 4: Distribution of the mixture fraction, reaction-progress variables and DMP on the outow surface for Rej = 400 and tr = 4.8 ms. Top row: short outlet tube (L/d = 1.62). Bottom row: long outlet tube (L/d = 10).

The dependence of the conversion of DMP on the inlet Reynolds number and the reaction time scale is shown in Fig. 5. As expected, the conversion decreases when Rej increases, indicating that poor mixing favors the slow reaction. The conversion increases when tr decreases, as a result of the higher reaction rate. The computational results and the experimental measurements are in excellent agreement for Rej 400 and tr 9.5 ms. The discrepancy for Rej < 400 is not a surprise if we recall the limitations of the turbulence model at low Reynolds numbers. When the reaction involving DMP becomes much faster (for example, tr = 4.8 ms), diculty in the experimental measurements has 8

Ying Liu, Qing Tang, and Rodney O. Fox


100

+ +

+ + + + + + ++ + ++

10-1

10-2

10

-3

317ms 181ms 61ms 28ms 16.7ms 9.5ms 6.5ms 4.8ms 317ms, sim 181ms, sim 61ms, sim 28ms, sim 16.7ms, sim 9.5ms, sim 6.5ms, sim 4.8ms, sim X max, sim

++ + ++ + ++ + ++ + + + + + + + + + + + +

X
101

102

103

104

Rej
Figure 5: Conversion of DMP versus Rej in the CIJR. Open symbols: experiments. Closed symbols: simulations.

been reported 11 . Additionally, Fig. 4 indicates that DMP in the sample collected during the experiments continues to convert in the outlet tube, especially for small tr . Thus, we have reasons not to blame the CFD models for the discrepancy observed in Fig. 5 12 . Overall, the ability of the CFD model to predict mixing-sensitive reactions in liquidphase systems is found to be satisfactory provided, of course, that accurate expressions are available for the chemical kinetics. 4 NON-PREMIXED COMBUSTION

Modeling non-premixed turbulent combustion is a great challenge because ames normally exhibit dierent levels of nite-rate chemistry eects ranging from near equilibrium to near global extinction. The amelet model or conditional moment closure based on a conserved scalar (i.e. the mixture fraction) is known to be quite reliable for the combustion processes characterized by fast chemistry 4. However, they fail in the slow-chemistry regime. Transported PDF methods are found to be able to predict the experimental results of combustion accurately 9. Unfortunately, the prohibitive computational cost limits their applicability when modeling practical combustion devices. Thus, there is still room for intermediate approaches with the ability to model complex ames accurately, but at a lower cost than transported PDF methods. Here, the application of a two-environment DQMOM-IEM model in modeling niterate combustion will be briey explored. Masri and coworkers 13 have experimentally investigated a blu-body burner that bears a great similarity to practical combustors. Non-premixed streams of air and fuel with a 1:1 (volume) mixture of methane and hydrogen enter the burner. In the DQMOM-IEM model (Eq. 2), the air stream corresponds to environment 1 while the fuel corresponds to environment 2. The ow rate of the air stream Uair is kept at 40 m/s while the inlet ow rate of the fuel jet Ufuel is 118 m/s 9

Ying Liu, Qing Tang, and Rodney O. Fox

x/DB=0.26
2500

x/DB=0.9
2500
2500

x/DB=1.3

2000

2000

2000

mean T

mean T

mean T

1500

1500

1500

1000

1000

1000

500

500

500

0 0 0.5 1

0 0 0.5 1

0 0 0.5 1

2500

y/Rb
2500

y/Rb
2500

y/Rb

2000
2000
2000

mean T

mean T

mean T

1500

1500

1500

1000

1000

1000

500

500

500

0 0 0.5 1

0 0 0.5 1

0 0 0.5 1

y/Rb

y/Rb

y/Rb

Figure 6: Radial proles of the mean temperature. circles: experiments, , simulations for ame HM1; +: experiments, : simulations for ame HM3. Top row: transported PDF; bottom row: DQMOM-IEM.

(ame HM1) or 214 m/s (ame HM3). Given the turbulent ow statistics from a CFD turbulence model, the DQMOM-IEM model can be used to treat the micromixing and chemistry. Here the augmented reduced mechanism 14 consisting of 12 reactions and 19 species is used to model the chemistry. Taking the enthalpy in each environment into consideration, the number of variables to be modeled by the DQMOM-IEM model is 2 20, in addition to the probability of environment 1. By adopting a time-splitting scheme, Eq. 2 is decomposed into a partial dierential equation (PDE) that involves the convection, diusion, micromixing and correction term and a ordinary dierential equation (ODE), the right-hand side of which is nothing but the chemical source term. Then the ODE can be solved using the in-situ adaptive tabulation (ISAT) algorithm 15 , which is favored for its high computational eciency. More related numerical strategies can be found elsewhere 16. The bottom row of Figs. 68 shows the calculated radial proles of mean temperature, and the mass fractions of CO and NO, respectively, at three axial locations. Predictions by a stand-alone joint velocity-turbulence frequency-composition PDF method 17 are displayed in the top row for comparison purposes. For ame HM1, the transported PDF approach and the DQMOM-IEM model predict the radial distributions of the mean temperature, mass fractions of CO accurately while both overpredict the mass fraction of NO at x/DB 0.9, y/Rb 1.0. For ame HM3, the DQMOM-IEM model does a better job in predicting the trends of the mean temperature, the mass fraction of CO and NO 10

Ying Liu, Qing Tang, and Rodney O. Fox

x/DB=0.26
0.1
0.1

x/DB=0.9
0.1

x/DB=1.3

0.08

0.08

0.08

mean CO

0.06

mean CO

mean CO

0.06

0.06

0.04

0.04

0.04

0.02

0.02

0.02

0.1

0.5 y/Rb

0.5 y/Rb

0.1

0.1

0.5 y/Rb

0.08

0.08

0.08

mean CO

mean CO

0.06

mean CO

0.06

0.06

0.04

0.04

0.04

0.02

0.02

0.02

0 0 0.5 y/Rb 1

0 0 0.5 y/Rb 1

0 0 0.5 y/Rb 1

Figure 7: Radial proles of the mass fraction of CO. circles: experiments, , simulations for ame HM1; +: experiments, : simulations for ame HM3. Top row: transported PDF; bottom row: DQMOM-IEM.

than does the transported PDF method. For example, the transported PDF calculations completely miss the trend of the mass fraction of CO at x/DB = 0.26 and the trend of the mass fraction of NO at x/DB = 0.9, while the DQMOM-IEM model successfully captures these trends. However, both of them tend to predict a mass fraction of NO higher than the experimental data. The CPU time consumed by the DQMOM-IEM model is only 3% of that consumed by the transported PDF method. We therefore conclude that for this example the DQMOM-IEM model provides an improved prediction but at a signicantly reduced computational cost. The same modeling approach can be used with LES 18 . Whether or not the DQMOM-IEM method will work satisfactorily for even more complex ames is still an open question. Nevertheless, its ease of implementation and the advantages it oers over simply neglecting the sub-grid scale uctuations makes it an attractive CFD tool for modeling practical combustion devices. 5 FINE-PARTICLE FORMATION

Other interesting examples of chemical reactors that can be modeled using CFD are those that produce ne particles. In many cases the ne particles are the desired product (e.g. nanoparticle formation in ames or colloidal particles), but in other cases they may be an undesired by-product (e.g. soot formation in ames.) Due to their small size (i.e. less than 10 microns in diameter), ne particles can often be treated as a pseudo species 19,20,21 11

Ying Liu, Qing Tang, and Rodney O. Fox

x/DB=0.26
0.015
0.015

x/DB=0.9
0.015

x/DB=1.3

0.01

0.01

0.01

mean NO

mean NO

mean NO
0.005

0.005

0.005

0
0 0.5 y/Rb 1

0.015

0.5 y/Rb

0.015

0.015

0.5 y/Rb

0.01

0.01

0.01

mean NO

mean NO

mean NO
0.005
0.005

0.005

0 0 0.5 y/Rb 1

0 0 0.5 y/Rb 1

0.5 y/Rb

Figure 8: Radial proles of the mass fraction of NO. circles: experiments, , simulations for ame HM1; +: experiments, : simulations for ame HM3. Top row: transported PDF; bottom row: DQMOM-IEM.

that follow the local uid velocity. Thus, in principle, any CFD model for turbulent reacting ows 4 could be used to describe their formation. However, the physical and chemical processes that lead to transformations in the properties of ne particles are rather complicated and lead to new challenges that are not present in the two reacting ows discussed earlier (i.e. mixing-sensitive chemistry and diusion ames.) In order to describe a population of ne particles, we must introduce the number density function (NDF) n(v ) representing the number density of particles with volumes in the interval (v, v +dv ). Note that for each value of v , n(v ) depends of the spatial location and time. Thus, it is a pseudo chemical species, and in fact the transport problem involves an innite number of such species parameterized by v . From the standpoint of CFD, we are therefore dealing with a turbulent reacting ow with an innite number of reacting scalars. Note that in most practical applications such a nanoparticle production in ames, we must include the normal reacting scalars associated with the chemistry and energy balance (denoted by ), and that these scalars are strongly coupled to n(v ). The rst step in developing a CFD model for ne-particle formation in turbulent ow is to write down the microscopic transport equation (i.e. the laminar ow model) for the scalars. For the standard scalars this is straightforward and can be found in textbooks. For the NDF, the transport physics (due to nite-size eects and particleparticle interactions) are more complicated. Nevertheless, for discussion purposes, let us 12

Ying Liu, Qing Tang, and Rodney O. Fox

assume that the microscopic transport equation has the form n + (U n) = [(, v )n] + Sn (, v ). t (5)

In this equation, U is the local uid velocity (assumed to be the same as the particle velocity, (, v ) is the size- and (possibly) composition-dependent molecular diusivity of particles of size v , and Sn (, v ) is a complex source term that models the changes in n(v ) due to nucleation, growth, aggregation, breakage, etc. We should note that this source term is typically an integro-dierential equation that by itself is dicult to solve numerically. In fact, there are a wide range of methods for treating the homogeneous problem: dn = Sn ( , v ) , (6) dt such as Monte-Carlo simulations, sectional methods, and moment methods. However, when applied to inhomogeneous problems (i.e. solving the full transport equation given in Eq. 5), only moment methods are currently tractable. The application of moment methods to treat Eq. 5 starts by dening the moments of the NDF: v k n(v ) dv. (7) mk =
0

Applying this transformation to Eq. 5 leads to mk + (U mk ) = J k () + Sk () t where the diusive ux of the k th moment is dened by

(8)

J k ( ) =
0

(, v )(v k n) dv,

(9)

and Sk () is the source term for the moments. Note that (even for a laminar ow) neither J k nor Sk will normally be closed terms (i.e. dependent on only the moments mk .) Thus, moment methods lead to closure problems in even the simplest ows. A powerful approach for closing the moment equations is the quadrature method of moments (QMOM) 22, which expresses the NDF moments in terms of a nite set of weights and abscissas:
M

mk =
m=1

k wm vm .

(10)

In practice, excellent predictions are often possible with small values of M (25). Thus, the innite set of scalars n(v ) is replaced by a total of 2M scalars (i.e. a set of 2M NDF moments) whose microscopic transport equation is Eq. 8. In practice, QMOM is feasible numerically due to the existence of the product-dierence (PD) algorithm that solves 13

Ying Liu, Qing Tang, and Rodney O. Fox

Eq. 10 for wm and vm given the set of NDF moments M = (m0 , m1, . . . , m2M 1). In other words, if you know M (by solving Eq. 8), then you can rapidly compute the corresponding weights and abscissas. The latter are used to close the diusive ux:
M

J k ( )
m=1

k (, vm)(wm vm ).

(11)

and the source term Sk (). Since the PD algorithm can not be extended to more than one variable, Marchisio & Fox 23 developed the DQMOM approach to derive the transport equations for the weights and abscissas directly (i.e. instead of solving the transport equations for the NDF moments). Nevertheless, for a uni-variate NDF using either QMOM or DQMOM is entirely equivalent. Thus, to summarize what we have discussed so far, using QMOM it is possible to treat ne-particle formation in laminar ows or in turbulent ows using direct numerical simulations (DNS) with a nite number of scalars. However, the problem of treating ne-particle formation in turbulent ow using a CFD model (e.g. RANS or LES) still remains to be discussed. If you consider the complete set of scalars describing the chemistry and the ne particles ( and M), we can see that conceptually the modeling problem is the same as treating any turbulent reacting ow with complex chemistry 4. For example, the treatment of turbulent reacting ows can be done by combining the multi-environment DQMOM-IEM model with the transport equations for the NDF moments using QMOM to nd the weights and abscissas. Alternatively, we can combine the multi-environment DQMOM-IEM model with the transport equations for the weights and abscissas directly (i.e. DQMOM). In these representations, the nth environment contains a set of NDF moments mkn or, equivalently, a set of weights and abscissas (wmn , vmn ). The CFD model equations for the weights and abscissas in each environment, which are very similar to Eq. 2, are obtained from Eq. 8 and have the form 24 Dpn = (T pn ) , (12) Dt Dpn wmn = [T (pn wmn )] + b (13) n (wm ) , Dt Dpn wmn vmn = [T (pn wmn vmn )] + c (14) n (wm vm ) . Dt In these equations, b n and cn are the combined terms for micromixing and the NDF moment source terms. These terms can be found for each environment (n = 1, , Ne ) by solving the linear system dened by 23
M M k vmn b n m=1

(1 k )

(wm ) + k
m=1 M

k 1 vmn cn (wm vm) = bn (mk ) + pn Mkn + pn Skn

+ k (k 1)
m=1

k 2 pn wmn vmn T |vmn |2

for k = 0, 1, . . . , 2M 1; (15)

14

Ying Liu, Qing Tang, and Rodney O. Fox

where bn , Mkn and Skn denote the closed terms for correction, micromixing and the sources due to nucleation, growth, aggregation and breakage for the k th-order NDF moment in the nth environment, the transport equation of which is in the form of Eq. 2: Dpn mkn = [T (pn mkn )] + pn Mkn + pn Skn + bn (mk ) . Dt (16)

In the example given below, we will consider only aggregation. More details on the extension to multivariate population balances can be found elsewhere 24. The key conceptual idea when thinking about the CFD model for ne-particle formation in turbulent ow is that each environment (or uid particle) has its own NDF, and thus its own set of NDF moments (and weights and abscissas). Note that this is entirely consistent with the usual one-point statistical description of turbulent reacting ows 4 . Thus, for example, in a RANS model we will have a Reynolds-average NDF denoted by n(v ) and Reynolds-average NDF moments denoted by mk . In the context of multienvironment models, the latter are computed in the usual manner:
Ne

mk =
n=1

pn mkn .

(17)

Note that this also implies that the NDF moments (like any other one-point scalar eld) will have Reynolds-average moments of arbitrary order. For example, the second-order moment of mk is
Ne

m2 k

=
n=1

pn m2 kn .

(18)

The existence of both the NDF and the one-point PDF of the NDF (!) can be a source for confusion (especially when stochastic solution methods are used to nd approximate solutions for the NDF). The reader should keep in mind that the NDF is not a probabilistic quantity, but instead arises due to the innite possible number of particle sizes present in the system. In contrast, the one-point PDF arises from the statistical modeling approach used to describe turbulent mixing. As mentioned above, QMOM and DQMOM should yield identical results for the statistics of the NDF moments mk . However, from a numerical perspective it is advantageous to use DQMOM. For example, when discretizing the CFD model for the NDF moments in each environment mkn numerical errors (primarily in the convection term) can lead to realizability problems (e.g. m2 1n > m2n m0n ). Such errors are avoided with DQMOM if the numerical scheme is chosen such that the weights and abscissas are everywhere non-negative. As an example, we consider the application of a two-environment (Ne = 2), four-node (M = 4) CFD model in a 1-D periodic domain 25 that satises the following assumptions: (i) the turbulence eld is homogeneous and stationary; (ii) the mean velocity is zero everywhere; (iii) the dimensionless length of the domain is two and the 15

Ying Liu, Qing Tang, and Rodney O. Fox


1

0.8

0.6

p1
0.4 0.2 0 0

0.5

1 X

1.5

Figure 9: Initial distribution of the inlet streams in the x-direction.

dimensionless turbulent diusivity is 0.1286. Initially the domain is non-premixed and separated at x = 1 into two parts (Fig. 9). One part contains only environment 1 in which w11 = 0.997, w21 = w31 = w41 = 0.001, v11 = 1, v21 = 2, v31 = 3, and v41 = 4. The other part contains only environment 2 in which w12 = 0.9985, w22 = w32 = w42 = 0.0005, v12 = 0.5, v22 = 1.5, v32 = 2.5, and v42 = 3.5. The initial distribution of p1 across the domain is shown in Fig. 9. As time evolves, the weights and abscissas of each environment change due to mixing and aggregation (Brownian aggregation in this example). Using either QMOM or DQMOM, eight NDF moments are needed for each environment. In Fig. 10 four NDF moments are compared for the two methods. In this example, the rst two terms on the right-hand side of Eq. 15 were neglected when we calculated b n and cn for numerical stability purposes. Any dierences between the QMOM and DQMOM results are due to neglecting these terms. Algorithms for treating those two terms stably and accurately are under development. As Fig. 10 shows, NDF moments predicted by DQMOM and QMOM are in close agreement. The distributions of the zero-order NDF moment in the two environments (m0n ) are identical since the initial values of them are identical and aggregation rates are comparable in both environments. The zero-order NDF moments decrease gradually as a result of the decreasing number density due to aggregation. The rst-order NDF moment in environment 1 decreases with time, while in environment 2 it increases. Although it is not shown in the gure, the Reynolds-averaged rst-order NDF moment ( m1 = p1 m11 + p2 m12) remains constant, indicating that neither mixing nor aggregation changes the total volume of the particles in this closed system. The higher-order NDF moments increase monotonically since the abscissas increase due to aggregation. Nevertheless, the third-order NDF moments increase faster than the second-order NDF moments. Due to micromixing, the NDF moments in the environments approach the Reynolds-average NDF moments at large times. In the example presented above, we considered only aggregation and mixing (i.e. mesomixing due to turbulent dispersion and micromixing). In more challenging applications such as soot production 26 there will be a strong coupling between the chemistry (i.e. ) and the evolution of the NDF. For example, soot will be formed at locations in composi16

Ying Liu, Qing Tang, and Rodney O. Fox

Figure 10: Time evolution of the spatial distribution of the moments mkn (k = 0, . . ., 3) predicted by DQMOM (lines) and QMOM (symbols). Left: Environment 1. Right: Environment 2. Square: initial conditions. Down triangle: t = 0.002. Circle: t = 0.02. Diamond: t = 0.04. Up triangle: t = 0.08.

17

Ying Liu, Qing Tang, and Rodney O. Fox

tion space that are fuel rich and will be oxidized in locations with excess oxygen. It can be anticipated that the successful description of ne-particle formation is such strongly coupled problems will require a detailed CFD model that explicitly accounts for subgrid-scale uctuations and the correlations between and the NDF. 6 CONCLUSIONS

As illustrated in this work, CFD models for single-phase turbulent mixing in chemical reactors are quite advanced and are now available in many commercial CFD codes. As highlighted in the examples, the basic components for successful predictions are accurate models for 1. the turbulent ow eld, 2. turbulent scalar transport (dispersion and advection), 3. subgrid-scale uctuations in the scalar elds (micromixing), 4. chemical/physical processes such as reactions and particle dynamics. If at all possible, each of these model components should be validated independently. In any case, the quality of the overall CFD predictions depend on the accuracy of all component so each of them should be considered carefully. The extension of the CFD models discussed here to multiphase chemical reactors will require the addition of more physics. For example, if instead of ne particles the dispersed phase consists of large particles then it no longer suces to assume that the particle velocity is the same as the uid velocity. A CFD model for multiphase ows that considers a separate momentum equations for each phase can be used for this purpose. In addition, it will also be important to include accurate models for heat and mass transfer between phases, and chemical reactions inside the phases or at the phase boundaries. As noted in the Introduction, each new reactor type will require an appropriate CFD model that captures the ratecontrolling steps in the overall process. Thus, the eld of CFD modeling of chemical reactors still has many new horizons to explore. ACKNOWLEDGMENTS We gratefully acknowledge nancial support from the U.S. National Science Foundation through a regular grant (CTS-0336435) and the SBIR Phase I program (DMI-0441833). We would also like to acknowledge our many collaborators with whom we have worked to carry out the research used in the example applications discussed in this work. References [1] R.V. Mehta, J.M. Tarbell. An experimental study of the eect of turbulent mixing on the selectivity of competing reactions, AIChE Journal, 33, 10891101, (1987). 18

Ying Liu, Qing Tang, and Rodney O. Fox

[2] J.A. Baldyga, R. Pohorecki. Turbulent micromixing in chemical reactors a review, Chemical Engineering Journal, 58, 183195, (1995), [3] J.A. Baldyga, J.R. Bourne. Turbulent mixing and chemical reactions. New York: John Wiley & Sons, (1999). [4] R.O. Fox. Computational Models for Turbulent Reacting Flows, Cambridge University Press, (2003). [5] H. Feng, M.G. Olsen, Y. Liu, R.O. Fox, J.C. Hill. Investigation of turbulent mixing in a conned planar-jet reactor, AIChE Journal, 51, 26492664, (2005). [6] H.C. Chen, V.C. Patel. Near-wall turbulence models for complex ows including separation, AIAA Journal, 26, 641648, (1988). [7] U. Piomelli, J. Ferziger, P. Moin, J. Kim (1989). New approximate boundary conditions for large eddy simulations of wall-bounded ows. Physics of Fluids A , 1, (6), 1061-1068. [8] V. Raman, R.O. Fox, A.D. Harvey, D.H. West. Eect of feedstream conguration on gas-phase chlorination reactor performance. Industrial & Engineering Chemistry Research, 42, 25442557, (2003). [9] V. Raman, R.O. Fox, A.D. Harvey. Hybrid nite-volume/transported PDF simulations of a partially premixed methane-air ame. Combustion and Flame, 136, 327 350, (2004). [10] L. Wang, R.O. Fox. Comparison of micromixing models for CFD simulation of nanoparticle formation. AIChE J. 2004, 50, 22172232, (2004). [11] B.K. Johnson, R.K. Prudhomme. Chemical processing and micromixing in conned impinging jets. AIChE Journal, 49, 22642282, (2003). [12] Y. Liu, R.O. Fox. Predictions for chemical processing in a conned impinging-jets reactor. AIChE Journal, 52, 731744, (2006). [13] A.R. Masri. http://www.mech.eng.usyd.edu.au/research/energy/#data. [14] C.J. Sung, C.K. Law, J-Y Chen. Augmented reduced mechanisms for NO emission in methane oxidation, Combustion and Flame, 125, 906919, (2001). [15] S.B. Pope. Computationally ecient implementation of combustion chemistry using in situ adaptive tabulation, Combustion Theory and Modelling, 1, 4163, (1997).

19

Ying Liu, Qing Tang, and Rodney O. Fox

[16] Q. Tang, W. Zhao, M. Bockelie, R.O. Fox. Numerical simulations of turbulent blubody ames using multi-environment presumed PDF method with realistic chemistry, Fall Meeting of the Western States Section of the Combustion Institute, Stanford University, Paper 05F-36, (2005). [17] K. Liu, S.B. Pope. Calculations of blu-body stabilized ames using a joint probability density function model with detailed chemistry, Combustion and Flame, 141, 89117, (2005). [18] V. Raman, H. Pitsch, R.O. Fox. Eulerian transported probability density function sub-lter model for large-eddy simulations of turbulent combustion, Combustion Theory and Modelling, , (in press), (2006). [19] D. Piton, R.O. Fox, B. Marcant. Simulation of ne particle formation by precipitation using computational uid dynamics, Canadian Journal of Chemical Engineering, 78, 983993, (2000). [20] T. Johannessen, S.E. Pratsinis, H. Livbjerg. Computational uid particle dynamics for the ame synthesis of alumina particles, Chemical Engineering Science, 55, 177 191, (2000). [21] T. Johannessen, S.E. Pratsinis, H. Livbjerg. Computational analysis of coagulation and coalescence in the ame synthesis of titania particles, Power Technology, 118, 242250, (2001). [22] R. McGraw. Description of aerosol dynamics by the quadrature method of moments. Aerosol Science and Technology, 27, 255265 (1997). [23] D.L. Marchisio, R.O Fox. Solution of population balance equations using the direct quadrature method of moments, Journal of Aerosol Science, 36, 4373, (2005). [24] R.O. Fox. CFD models for analysis and design of chemical reactors. Advances in Chemical Engineering, in press, (2006) [25] L. Wang, R.O. Fox. Application of in situ adaptive tabulation to CFD simulation of nano-particle formation by reactive precipitation, Chemical Engineering Science, 58, 43874401, (2003). [26] A. Zucca, D.L. Marchisio, A.A. Barresi, R.O. Fox. Implementation of the population balance equation in CFD codes for modelling soot formation in turbulent ames, Chemical Engineering Science, 61, 8795, (2006).

20

You might also like