Chemical
Engineering
Science, Vol. 49, No. 248, pp. S229mS241, 1994 Copyright fQ 1995 Elsevier Science Ltd Printed in Great Britain. All rights reserved OW-2509194 57.C0+O.CQ
OCKW-2509(94)00299-l
REACTIVE MIXING IN A TUBULAR JET REACTOR: A COMPARISON OF PDF SIMULATIONS WITH EXPERIMENTAL DATA MASSIMO Dipartimento
PIPINO
di Scienza de1 Materiali e Ingegneria Chimica, Politecnico di Torino, C. so Duca degli Abruzzi 24, 10124 Torino, Italy
and RODNEY
0.
FOX*
College of Engineering, Kansas State University, Manhattan, (Accepted for publication
26 September
KS 66506, U.S.A.
1994)
Abstract-The simulation of turbulent reactive flows in chemical reactors can be extremely useful for practical applications. In particular, both industrial and academic chemical reaction engineers are keenly interested in the exact description of the composition field in a reactor volume. Results from two different probability density function (PDF) descriptions of a laboratory tubular jet reactor are compared with experimental data for acid-base neutralization in turbulent liquid media. The first description is based on the Lagrangian joint PDF of velocity and composition, while the second is based on the Eulerian composition PDF. In both descriptions chemical reactions are treated exactly, but molecular mixing must be modeled. A computational fluid dynamics code (FLUENT) provides the mean velocity field and turbulence quantities. Results are found to be sensitive both to turbulent diffusivity and to the model for molecular mixing. A new model for the effect of scalar integral-length scale relaxation on the molecular mixing rate is proposed. By comparison with experimental data, the superior performance of a new model over existing molecular mixing models is demonstrated.
1. INTRODUCTION
The simulation of turbulent reactive flows in real reactors can be extremely useful for practical applications. In particular, both industrial and academic chemical engineering researchers are keenly interested in predicting the composition fields in chemical reactors. The conventional deterministic approach to this problem based on solving moment equations for the composition variables is of limited use due to the well-known difficulties in closing the nonlinear reaction rate terms in turbulent flows. The probability density function (PDF) approach overcomes these difficulties by describing each point in the turbulent flow by an n-dimensional joint PDF of random variables such as the velocity and/or the chemical compositions. Depending on the level of closure, the joint PDF may include the fluctuating velocity components, thereby avoiding the closure problem (i.e. the gradient-diffusion assumption) associated with the composition-velocity fluctuation transport term in the composition moment balance equation. However, the most important advantage of the PDF approach for reactive flow applications is the exact treatment of the reaction term. PDF methods are thus a promising tool for chemical reaction engineering applications. *To whom correspondence should be addressed.
This work investigates two different PDF descriptions of a laboratory tubular reactor, comparing the predictions to experimental data for acid-base neutralization: A+BdP
(1)
with k, = 1.0 x 10’ ma/km01 s. The first description is based on the joint velocity-composition PDF transport equation for Lagrangian fluid particles traveling through the reactor (Pope, 1985). Each notional particle in the resulting Monte-Carlo simulation carries with it random variables specifying its location, velocity, and chemical composition. The turbulent motion is modeled by a stochastic differential equation whose coefficients are functions of the local mean velocity, the local turbulent kinetic energy, and the local turbulent energy dissipation (Pope, 1985; Tsai and Fox, 1994). Unlike in standard moment closures, the transport of chemical species due to velocity fluctuations is treated exactly at this level of description. However, the molecular mixing term must be closed, using a local characteristic mixing time found from the turbulence fields. As shown in the sequel, the proper specification of the molecular mixing rate is crucial for successfully modeling the experimental data. The second description is based on the transport equation of the Eulerian joint composition PDF. For
5229
MASSIMO PIPINO and RODNEY 0. Fox
5230
I
HCI
NaOH
Fig. 1. Experimental setup for tubular jet reactor this case, the velocity does not appear as a random
variable in the joint PDF, and thus a computational fluid dynamics code (FLUENT) is required to determine the mean velocity and the turbulent diffusivity. A novel Monte-Car10 solution algorithm has been developed that interfaces directly with FLUENT. The description, although in theory less rigorous than the previous one due to the need to model transport by turbulent velocity fluctuations, is nevertheless very promising for industrial applications since the lower dimensionality of the PDF allows for full 3D simulations or, in 2D cases, requires less memory and CPU time. In the present code, a gradient-diffusion approximation is employed. Using the PDF codes, simulations have been carried out to predict the composition fields in a laboratory tubular reactor with nonpremixed coaxial feeds. The extent of the reaction zone is determined experimentally by the use of a chemical indicator (phenol red) added to both feed streams. The axial concentration profile for the key reactant (fed in the inner tube) has been determined by a spectrophotometer with fiber-optic sensors able to move along the axis of the reactor in correspondence to the diametrical plane. Data have been collected for various Reynolds numbers, various inlet concentration ratios of the two reactants, and several internal tube geometries, and compared to simulation results from the two PDF codes. The remainder of this work is organized into five sections. Section 2 describes the experimental setup. Section 3 reviews the PDF approach and the models that close the PDF transport equation. A new model is introduced for the rate of molecular mixing which, unlike previous formulations, explicitly accounts for the relaxation of the concentration spectrum. Section 4 briefly describes the numerical procedures employed in the PDF simulations. Results from the PDF codes are compared with the experimental data in Section 5. Section 6 draws conclusions and discusses some possible extensions.
2. EXPERIMENTAL
SETUP
In order to check the results given by the simulations, axial concentration profiles have been measured experimentally in a tubular jet reactor with nonpremixed coaxial feed. This choice is due to the simple hydrodynamic field involved in such a geometry and to the fact that this configuration allows simple measurements of concentration decay. The experimental setup (see Fig. 1) was realized in the micromixing laboratory of Politecnico di Torino with the use of a silica tube (10 mm interval diameter, 1 m long) mounted on an optical bench, in which different coaxial feed tubes (0.7-1.2 mm internal diameter) could be inserted. Note that the outer-to-inner tube diameter ratio of approximately 10 implies that both radial mixing due to turbulent diffusivity and turbulent molecular mixing will strongly affect the measured profiles. The measurement of axial concentration profiles was carried out using a fiber-optic spectrophotometer, able to record absorbance along the reactor axis. Both nonreactive and reactive experiments were carried out. Results from the former were used to check the rate of radial mixing, and from the latter to test the molecular mixing models. In the nonreactive case, methylene blue was used as an inert tracer, and the axial absorbance profile was measured using visible light (658 nm). For the reactive cases, NaOH-HCI neutralization was employed, and the residual base concentration visualized by relating it to the absorbance of phenol red measured at 558 nm. With the present experimental setup, soda (001 N) could be fed into the internal stream at various velocity ratios (mean internal velocity/mean external velocity) ranging from l-10 and various concentration ratios (acid concentration: base concentration) ranging from 0.5-20. In order to allow the comparison of absorbance data with the “absorbance” field produced by the simulation codes, the linearity of the Beer-Lambert
5231
Reactive mixing in a tubular jet reactor law in the experimental concentration range was verified, together with the influence of the local pH on the change of color of the indicator. In the simulation codes, the local absorbance field was computed from the local mixture fraction PDF using the method described by Li and Toor (1986). 3. PDF
METHODS
Excellent introductions to PDF methods can be found in O’Brien (1980) and Pope (1985), and a summary of their application to chemical reaction engineering problems can be found in Tsai and Fox (1994) where the PDF transport equation is derived and models for closing the velocity-composition joint PDF are reviewed. Here, discussion will be limited to a brief review of the relevant PDF transport equations and the closure models employed in the PDF codes. In particular, the molecular mixing model will be discussed in detail due to its strong influence on the predicted axial concentration profiles for the reactive cases.
appears in a similar form in the composition joint PDF transport equation and is discussed in detail below. 3.2. Composition joint PDF The Eulerian composition joint PDF approach is based on the solution of the PDF transport equation for &(JI; x, t):
3.1. Velocity-composition joint PDF The transport equation for the velocity-composition joint PDF f(u, C; x, t) can be written for a constant density flow as
The terms on the left-hand side of eq. (6) represent, respectively, the change in time of the PDF, convective transport in physical space due to the mean velocity field and convective transport in composition space due to chemical reactions. These terms appear in closed form. On the right-hand side of eq. (6), diffusive transport in composition space due to molecular mixing, and transport in physical space due to turbulent velocity fluctuations, can be identified. They appear in terms of conditional expectations and must be modeled. The modeling of the molecular mixing terms is common to both the composition joint PDF and the velocitycomposition joint PDF approach and is discussed below. Compared to the velocity-composition PDF approach, the composition PDF presents the disadvantage of modeling turbulent transport in physical space. In this work, turbulent transport is assumed to follow the gradient-diffusion model:
where the scalar flux JF and the viscous force zij have the following forms:
(7) where I-r represents the turbulent diffusivity. When the k-e model is used to describe the turbulent motion field, we can express the turbulent diffusivity as
In eq. (2), the mean pressure ( p) can be obtained from the Poisson equation: V’(P)
=
d2(UiUj> --P
Z’
(5)
A remarkable feature of the PDF approach can be seen by observing that all the terms on the left-hand side of eq. (2) are closed, which include the turbulent convection, the production by the mean pressure gradient, gravity and chemical reactions. However, the three conditional expectation terms on the right-hand side, which represent transport in velocity space by viscous forces, and the fluctuating pressure gradient and transport in composition space by diffusive molecular mixing, must be closed. The first two of these terms involve models that are independent of the composition fields and are discussed by Pope (1985) and Tsai and Fox (1994). The final term
OTC
where C,, is an empirical constant (set here to 0.09) and bT is the turbulent Schmidt number. The standard value of bT is 0.7; however, comparisons of the nonreactive data with simulation results suggest that this value is too small (radial mixing is too fast). The gradient-diffusion assumption can fail in variable-density reactive flows, but it is accurate in representing many constant density flows. The effect of varying CJ=has been found to be significant for the tubular jet reactor used in this work. Good agreement with experimental results for the nonreactive case has been found using CT== 1.0. This value has thus been employed in all isokinetic jet simulations using the composition PDF code reported in this work. For the non-isokinetic jet simulations, a low Reynolds number correction (M. Pipino, 1994, personal communication) shown in Fig. 2, has been used in some cases. In the
5232
MASSIMO PIPINO and RODNEY 0.
Fox
constant density with no reaction gives:
we see that the molecular mixing term is modeled as
For nonhomogeneous conditions, the same model is used but with Q and (4.) replaced by their local values. Fig. 2. Low Reynolds number correction factor for turbulent diffusivity employed in nonisokineticjet simulations.
joint velocity-composition code br = 1.0 corresponds to a value of C,, in the range of 8-9 [standard value is 2.1 (S. B. Pope, 1993, personal communication)]. All simulations using the joint velocity-composition code reported in this work have been found with ca = 8. 3.3. Molecular mixing A successful model for closing the molecular mixing term must address two problems: first, a passive scalar PDF should relax to a Gaussian form regardless of its initial condition; second, the effect of turbulent stretching on the rate of scalar mixing should be properly accounted for. Several models have been proposed to address the first problem (Curl, 1963; Dopazo, 1975; Janicka et al., 1979; Pope, 1982, Gao, 1991; Valifio and Dopazo, 1991; Fox, 1992, 1994; Norris, 1993; Fox and Tsai, 1994). In this work, the LMSE (linear mean square error) mode1 will be employed. This model is exact when the initial PDF is Gaussian (O’Brien, 1980) and has the simplest mathematical form. For the simple acid-base reaction under consideration, the results are insensitive to the PDF shape; hence, the problem of turbulent stretching is more significant. In this section, a new model for the molecular mixing rate is proposed to explicitly account for relaxation of the concentration spectrum to its isotropic form through the actions of turbulent mixing. 3.3.1. LMSE mixing model. The LMSE model in isotropic, homogeneous turbulence with an isotropic scalar field reads
!&_at
-;
c&I - <&>I,
where r,+is the molecular mixing time scale determined by turbulent stretching. The corresponding PDF evolution equation can be written as
By comparing eq. (10) with the evolution equation of &, [Eq. (6)] under conditions of homogeneity and
3.3.2. Molecular mixing rate. For modeling the effect of molecular mixing on nonpremixed chemical reactions, the choice of the molecular mixing rate is paramount. Several models have been suggested in the literature to relate z+ to local turbulence quantities such as k and E. Baldyga and Bourne (1989) proposed a simplified micromixing model that incorporates the “hydrodynamic lifetime of a vortex” that they compute to be proportional to (v/s)‘/‘, e.g. the Kolmogorov time scale (Baldyga and Bourne, 1984). Using their model, the molecular mixing time can be written as 54.1 = 17.31
0 1
E
l/2
.
(13)
Due to the choice of turbulence time scale, this model is applicable when the rate of mixing is controlled by vortex engulfment at fine scales. The standard model for high Reynolds number isotropic turbulent flows relates the molecular mixing time to the eddy turnover time which is proportional to k/E (Pope, 1985): 2k
(14)
Q.,2 = ,-,
where C, has a limiting value of 2 for high Reynolds number flows. This choice of r+ (at least in the asymptotic decay regime*) is supported by direct numerical simulations of a nonpremixed scalar field in isotropic, homogeneous turbulence (Eswaran and Pope, 1988). Use of r+,s is appropriate when scalar gradients are created by integral-scale eddies and the rate of transport from integral scales to small scales is the rate-controlling step. Use of r’o,2 also assumes that the scalar spectrum is fully developed at all scales. Equations (13) and (14) differ by a factor proportional to the turbulence Reynolds number defined by (Hinze, 1975): 60 Re,=-- 43&
k
Thus, as the turbulence Reynolds number increases, T+.i will decrease much faster than T&.~. If scalar
*That below.
is, at the end of the relaxation
period modeled
Reactive mixing in a tubular jet reactor
gradients are initially created at scales near the integral scale of the turbulence field as is the case for turbulent jets, at high Reynolds number, r9,2 will be the appropriate limiting molecular mixing time. On the other hand, if the scalar gradients are created on smaller scales, as performed by Baldyga et al. (1994) by using a separate feed tube, r+,t may become the limiting molecular mixing time. The spectral relaxation model presented below explicitly accounts for different initial scalar integral-length scales by introducing a cascade of scalar dissipation from integral scales, through the Kolmogorov scale, down to the Batchelor scale. Corrsin’s model (Corrsin, 1964), which combines r+. I and rb. z was also tested:* 70.3
for SC >> 1. Note number, eq. (16)
=
i
i
+
5
ln(Sc)
Y 0&
5233
3.3.3. Spectral relaxation model. A scalar field advected by an incompressible turbulent flow is governed by
W - = n-Q,
(17)
Dt
and the magnitude of its gradient by (see Appendix)
where xi = @/axi and eij = &,/axi is the strain rate tensor (repeated indices imply summation). Reynolds averaging of eq. (17) multiplied by 4’ yields (see Appendix)
r/2
(19)
(16)
that at high turbulence Reynolds approaches eq. (14) if C, = 4/3.’
Note also that Corrsin assumed that the scalar spectrum was fully developed when deriving t+, 3. Experience, in this work, with the above models suggests that none are entirely adequate for a tubular jet reactor. For example, C, in eq. (14) must be set smaller than the limiting value in order to achieve even qualitatively satisfactory agreement with experimental data. In confined flows, where reaction occurs near the bounding wall, E attains high values in the boundary layer (i.e. Re, decreases) resulting in an overprediction of the mixing rate (Tsai and Fox, 1994). Furthermore, even for free jets, all three models may overpredict mixing rates near the injection point due to the lack of initial integral-length scale information and the assumption that the scalar spectrum is fully deve1oped.t Indeed, even in stationary isotropic turbulence, if the initial integral-length scale of the scalar field is larger than the integral-length scale of the velocity field, rg approaches k/c only asymptotically from above (Eswaran and Pope, 1988; Kosaly, 1989; Jiang and O’Brien, 1991).6 These observations suggest that an improved molecular mixing model is needed that includes the effect of Re, and incorporates integral-length scale information in the form of a spectral relaxation period where r+ moves to its limiting value.
*The integral-length scale of the velocity field has been taken to be proportional to k3/‘/c (Hinze, 1975) and the proportionality constant fixed by forcing the original model [e?. (69) in C orrsin (1964)] to agree with ro.s at SC = 1. This differs from C, = 2 (valid for SC = 1.0) due to a dependence on SC (Corrsin, 1964). :In pure mixing studies, this problem has been avoided in the past by adjusting the data to eliminate the initial decay region at the injector (e.g. Brodkey, 1966). However, for reacting flows this region cannot be ignored since a considerable percentage of fast reactions can occur there (Baldyga et al., 1994). “These studies have also demonstrated that the initial scalar-to-velocity micro-length scale ratio is unimportant in determining scalar decay rates.
where the source term can be expressed as $2 Likewise, Reynolds Appendix):
=
a<4> -
-. axi averaging
of eq. (18)
(20) yields (see
The scalar dissipation can be defined as E, = f-xf2, and the molecular mixing rate as
such that r+ = l,l. Note that both E+ and r4 are random fields since they involve the fluctuating variable xJ2. side of eq. (21) The second term on the right-hand corresponds to turbulent stretching of the scalar gradient, and the first term to its molecular dissipation. The last two terms are production terms: the first due to the mean scalar gradient and the second to the mean velocity gradient. Fox (1994), based on studies of diffusion in random scalar fields and direct numerical simulation of scalar mixing in isotropic stationary turbulence, proposed the following models for the molecular dissipation and turbulent stretching terms:
and
r
= -G
0E “2 ;
(251
where C, = 0.54 (Fox et al., 1992; Leonard, 1989) is near Batchelor’s constant (Batchelor, 1952), and C, = 3 for one-dimensional diffusion in random
MAWMO PIPINOand RODNEY 0.
5234
Fox
2 E l/Z 3(azt3 Re, - 1) 0 Y
lamellar systems (Fox, 1994). Finally, if (xix>> is assumed to be nearly isotropic, the source term due to the mean velocity gradient can be neglected for incompressible flows.* Following Newman et al. (1981), the source term due to the mean scalar gradient can be modeled by
(29) 2
E 112
3(a 2’3 Rr, - 1) 0 ; s 112 2 <%) ln(Sc) (-1v
or (27)
0 E
where C, in the spectral relaxation model is uniquely determined in terms of C, and C, as discussed below. Note that the source term for (S,,/T) is proportional to .+ divided by the square of the Batchelor length scale, and thus represents a source for the maximum attainable scalar dissipation (i.e. with all scalar variance located at dissipative scales.) The source term differs from that employed by Mantel and Borghi (1991) where the Taylor length scale is used in place of the Batchelor length scale in order to obtain the correct limiting behaviour for . The spectral relaxation model describes the cascade of scalar dissipation from large to small scales by introducing additional variables that correspond to “potential” scalar dissipation residing at length scales larger than the Batchelor length scale. These variables are referred to as potential scalar dissipation because they are formed by dividing the scalar variance in a given range of length scales by the square of the Batchelor length scale. Thus they represent the maximum attainable scalar dissipation for a given scalar spectrum. The cascade model we propose is similar in spirit to the three-stage model proposed by Baldyga (1989) for the scalar variance. We denote the potential scalar dissipation ith stage of the cascade by E$. The cascade
in the
will be composed of up to three stages representing transport from the integral scale of the turbulence field, L,, to the initial7 integral scale of the scalar field, L,; transport from L, to the Kolmogorov scale of the turbulence field; and transport from the Kolmogorov scale to the Batchelor scale (Corrsin, 1964). Based on the form of eq. (21), we define the three-stage cascade model for SC r
1 and ReF3”
< c( = 2Lg/3L,
I
1 bys
(28)
*In Mantel and Borghi (1991), the source term is modeled by assuming it to be proportional to (uiu;> which is null in isotropic turbulence. +The “usual” initial condition is L+ = 2L,/3 (Corrsin, 1964). However, in some cases the scalar may be generated by a different mechanism than the turbulence field. For example, in direct numerical simulations the fields may be initialized independently of one another. Experimentally, L, << L, can be realized by using a separate feed tube for key reactants (Baldyga et al., 1994).
x +Cw where
Re,
=
=k. fi
-
V
l/Z
(33)
Note that eqs (19) and (23) are used to relate the scalar variance to the (E+). Comparison of the above model with direct numerical simulation data at Re, = 49, SC = 0.7 (two-stage model) and experimental data at Re, = 35, SC = 450, yield good agreement as discussed below. The source term is placed at the velocity integral scale since it describes gradient production due to the product of the scalar flux and the mean scalar gradient, both of which are integral-length scale terms. For the case where all scalar dissipation is produced by the source term (all dissipation variables are initially zero), a = 1 and stage 1 can be dropped. A further refinement of the model could include a dynamical equation for a in order to handle nonsteady flows. The dependence of the decay rates on Re, and SC follow from Corrsin (1964) and essentially represent the time needed for a peak in the scalar spectrum at one stage to arrive at the next stage. For small values of Re, and SC, these decay rates are large compared to the scalar dissipation rate, and stages 2 and 3 have little influence on the behavior of <.Q >. For this case, a one-stage model wherein eq. (29) decays directly into eq. (31) can be employed. For our reactor (Re, z 35, SC = 450), a one-stage model found by dropping eqns (28) and (30), taking a = 1 and Re,, SC >> 1 in eq. (29), and C, = 0.18 in eq. (31) gives good agreement with the full model.
*For 0 < I, the integral scale of the scalar field is above the integral scale of the turbulence field and stage 1 is dropped. For c( 5 ReF3” 9the integral scale of the scalar field is below the Kolmogorov scale.
Reactive mixing in a tubular jet reactor
Three first three terms on the right-hand side of the balance equation for [eq. (31)] model, respectively, the cascade (source) from larger scales, molecular dissipation, and turbulent stretching in the viscousdiffusive subrange of the scalar spectrum. The latter is taken directly from the model proposed by Fox (1994) [eq. (25)]. The characteristic dissipation rate in the molecular dissipation term has been modified [cf. eq. (24)] to include the sum of both the actual and potential scalar dissipation rates. This modeling assumption is justified for two reasons: (1) it yields the correct limiting behavior as Re, + 00, and (2) C, can be taken as independent of the initial scalar-tovelocity integral-length scale ratio and of Re,. The dependence on SC has been added to agree with Corrsin (1964) in the constant mean gradient case. Limiting cases are discussed below. Direct numerical simulation results (Fox et al., 1992) indicate that C, is independent of the initial scalar-to-velocity integrallength scale ratio, but that C, in eq. (24) exhibits a strong dependence on it. When this ratio is greater than unity, C, is initially very large and relaxes slowly to its final value. By including the source cascade (source) term and by modifying the molecular dissipation term, C, = 3, independent of SC, Re, and the initial scalar-to-velocity integral-length scale ratio.
5235
Fig. 3. The mechanical-to-scalar time-scale ratio, R, for isotropic, stationary turbulence (Re, = 49, SC = 0.7) found from the spectral relaxation model with t* = et/k. Solid line: (a = 0.989) = 0. Dashed line: (u = 0.456) = 8.25E/k, (~(0)) = 0.25&/k. Dash-dot line: (a = 0.300) = 0.75&/k. Asterisks: (a = 0.168) = 0.67&/k, = 11.25&/k, = 1.756/k. Dotted line: (a = 0.121) = 0.67&/k, = 1.0&/k. Crosses: (K = 0.121) = 0.67&/k, = 2.75&/k.
actual scalar dissipation quickly becomes constant: 3.3.4. Limiting cases. The limiting case of homogeneous turbulence and scalar fields with S,, constant (constant mean scalar gradient) can be used to relate C, to C, and C,. When Re, is very large, the scalar dissipation rate is controlled by the rate of transport from large to small scales so that (rd} + E/k (SC = 1). For SC >> 1, the limiting behavior has been determined by Corrsin (1964). Setting the material derivatives equal to zero in eqs (19) and (28)~(31), letting G(= 1, and taking the limit Re, + m, we find that c,=-.
CL0 c, - 1
(34)
For finite Re,, t,+ = k/e + 1.23(v/e)“’ when SC = 1. The limiting case of isotropic, stationary turbulence with no mean scalar gradient (S,l = 0) is also of interest. For this case, the material derivatives in eqs (29)-(31) become total derivatives and the behavior of _* In particular, at moderate to large Reynolds numbers, if