Sei sulla pagina 1di 18

Article pubs.acs.

org/IECR

A New Approach to Fixed Bed Radial Heat Transfer Modeling Using Velocity Fields from Computational Fluid Dynamics Simulations
Mohsen Behnam,, Anthony G. Dixon,*, Michiel Nijemeisland, and E. Hugh Stitt

Department of Chemical Engineering, Worcester Polytechnic Institute, 100 Institute Road, Worcester, Massachusetts 01609-2280, United States Johnson Matthey, P.O. Box 1, Belasis Avenue, Billingham, Cleveland TS23 1LB, U.K. ABSTRACT: A new velocity-based approach to xed bed radial heat transfer is presented. Axial and radial velocity components were averaged from detailed 3D computational uid dynamics (CFD) xed bed simulations of computer-generated beds of spheres and used to model radial thermal convection. The convection terms were coupled with a radially varying stagnant bed thermal conductivity in a 2D pseudocontinuum xed-bed heat transfer model. The usual eective radial thermal conductivity kr and apparent wall heat transfer coecient hw were not used, and there were no adjustable parameters. The radial and axial temperature variation predicted by the velocity-based model agreed well with the angular-averaged temperatures from the detailed 3D CFD simulations over the range 80 Re 1900 and for N = 3.96, 5.96, and 7.99.

1. INTRODUCTION Heat transfer in xed bed tubes is an important topic in the chemical industry because xed beds are extensively used in applications with heat eects, such as reactors, thermal storage units, and adsorption or desorption plants. In particular, multitubular xed bed reactors with low tube-to-particle diameter ratio (N) are used for extremely exothermic or endothermic reactions such as partial oxidations and steam reforming of methane, respectively. Heat must be rapidly transferred into or out of a narrow reactor tube, in which the tube wall has a strong inuence on heat transfer and ow of reactants around the catalyst particles. These in turn aect catalyst activity, selectivity, and deactivation. Current reactor models for heterogeneous gassolid reactors have been based on fairly radical simplifying assumptions, such as pseudohomogeneity, eective transport parameters, and uniform catalyst pellet surroundings. Despite the realization that local ow structures are critically important in determining the global behavior of a ow or transport system,1 in many cases the hydrodynamic modeling of reactors is still based on unidirectional axial plug ow. All mechanisms for radial heat transport are lumped into an eective radial thermal conductivity kr, which is taken as constant and used to describe heat transfer up to the wall. The observed increase in resistance to heat transfer near the containing wall has been a continuing source of diculty. The classical approach to modeling this increased resistance near the wall is to idealize it to occur at the wall, and lump all the mechanisms into a wall heat transfer coecient, hw. Thus the near-wall resistance is misplaced, and the temperature of the near-wall particles is under-predicted (for wall heating) along with the associated reaction rate. For narrow tubes this can be a major problem. A review has recently been presented of the present state of research and understanding of radial heat transfer in xed beds.2 The classical eective parameter kr hw model was extensively described and problems with typical approaches to obtaining and analyzing experimental heat transfer data to get kr and hw were
2013 American Chemical Society

explained. Current correlations for kr were evaluated, and the debate over the meaning and usefulness of hw was elaborated, in the context of the historical development of the concept. A discussion of alternatives to the kr hw approach and their pros and cons was made, with focus on recent eorts to include limited aspects of the velocity eld and local bed structure. The major nding from this review was that all current approaches to radial xed bed heat transfer modeling suer from serious deciencies, which motivates the completely new development of the present work. A new approach to modeling radial heat transfer in xed beds is proposed. The motivation for this approach is to discard the use of eective conduction to represent transport by convective motion of the uid, which is a purely uid mechanical phenomenon. A second motivation is to improve the prediction of temperature proles and heat uxes at the reactor tube wall by incorporating the physical phenomena that cause the extra resistance to heat transfer near the tube wall directly into the model, and not idealizing them at the wall with an articial temperature jump. We therefore postulate a new type of twodimensional pseudocontinuum model. For this work, the model is also pseudohomogeneous, but this is for convenience only and is not an essential part of the formulation. The development of this new approach depends on the use of 3D computational uid dynamics (CFD) for full beds of particles3,4 to obtain the necessary information on individual velocity components. The CFD simulation is then also used to provide validated temperature proles to test the 2D pseudocontinuum model. The strategy for this development is illustrated in Figure 1. The rst step is to develop 3D CFD discrete particle models for the detailed ow through a catalyst packing, to provide the detailed ow elds that are responsible
Special Issue: NASCRE 3 Received: Revised: Accepted: Published:
15244

January 12, 2013 March 15, 2013 March 20, 2013 March 20, 2013
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research

Article

Figure 1. Schematic of the relationships between the models investigated in this study.

for convective transport. Methods to generate the packings and mesh development strategies are an essential part of this step. The velocity and pressure elds then must be incorporated into a 3D energy balance to give a 3D temperature eld. This is then to be averaged to provide a 2D temperature eld T(r,z) which is the basis for comparison to our new 2D pseudocontinuum model. For this new model, the 3D velocity elds from the CFD model simulations have to be coarse-grained or averaged to extract the essential 2D information to be supplied to the pseudocontinuum model. Our early work5 suggested that we would need vz(r) and vr(r,z), to properly capture the radial convective transport of heat and the near-wall changes in transport. In addition, we use the well-established ZehnerSchlun der formula6 for conduction heat transfer in a stagnant bed to represent the conductive contribution ke0(r) to the overall heat ux. These contributions are both included in the pseudocontinuum model. The resulting 2D temperature elds are then compared against the 2D temperature elds obtained from direct averaging of the 3D CFD results. Further details of each of these steps in Figure 1 are provided in the following sections.

2. BACKGROUND TO VELOCITY-BASED MODELS The traditional approach to xed bed radial heat transfer for the last 60 years7,8 has been the classical two-dimensional pseudocontinuum heat transfer model and its boundary conditions as embodied in the following equations:
c pu0
kr

1 T T r = kr r r r z
= hw (T |r = R Tw )
r=R

(1)

T r

(2)

T r

=0
r=0

(3) (4)

T |z = 0 = Tin(r )

Several variations of this basic plug ow (PF) model exist, including those that incorporate axial dispersion or conduction terms. In this model the eective radial thermal conductivity kr was used to represent all mechanisms for radial heat transfer inside the bed, such as conduction, convective radial displacement of uid, and particleparticle radiation. Despite evidence
15245

that kr varied with tube radius, especially near the tube wall, it was usually taken as constant to simplify both parameter estimation and solution of the model equations. The strong decrease in kr as the tube wall was approached was idealized to be a heat transfer resistance located at the wall, and was represented by the wall heat transfer coecient, hw, along with a temperature jump at the wall. The parameters kr and hw each reect the eects of several dierent heat transfer mechanisms, and have proved dicult to determine over the years, especially at low N, while reaction models based on them have been criticized as being oversimplied. Several papers and reviews have addressed these concerns, many of which were summarized recently.2 One of the main perceived failings of the PF model has been the use of a radially uniform axial velocity u0 to represent ow in the tube. A comprehensive review of uid ow in packed tubes up to 1987 was given by Ziolkowska and Ziolkowski9 which demonstrated that the prevailing opinion was that the constant u0 should be replaced by vz(r) and the wall heat transfer coecient hw should not be used. Several research groups have developed various approaches to obtain vz(r), including extended capillary models,10 the extended BrinkmanDarcyForchheimer equation from either particle-based11,12 or porous-media based1315 methods, a combination of these two models,16 and various models derived from the volume-averaged NavierStokes equations.1720 Several of these ow models found it necessary to introduce an eective viscosity into the equations,12,17 which introduces another parameter that requires estimation and to some degree negates the advantage of dispensing with the wall coecient. In addition, the eective radial thermal conductivity is retained, but must now be re-estimated for use in the altered model.21,22 In parallel with the problems associated with heat transfer modeling in packed tubes, several authors have expressed dissatisfaction with the standard dispersion model (SDM) which uses eective diusion to represent axial and radial dispersion23 and has drawbacks including innite speed of propagation and overestimation of back-mixing. Some dierent approaches to this problem have included the cross-ow model,24 the alternating ow model,25 and the wave model, rst put forward by Stewart26 and more recently strongly championed by Kronberg and his colleagues.27,28 The application of the wave model to xed bed heat transfer was demonstrated by Kronberg and Westerterp27 whose work showed that this model also results in parameters that must be determined from experimental data. Kronberg and Westerterp27 in particular presented a strong argument for the need for a new approach to modeling transport in xed beds. The crux of their argument was that for many years we have used eective diusion and conduction models to represent heat and mass transfer phenomena that are essentially uid mechanical in nature. The reason for this has largely been computational convenience, a constraint that is rapidly being eased by the development of faster, larger computers and improved numerical methods. It should be possible to move toward models that more realistically represent the ow eld in a xed bed. Some of these points of view have more recently been echoed by Schnitzlein29 who pointed out that eective dispersion coecients are commonly used with gradients in concentration and that a large contribution to dispersion is uid mechanical which is solely driven by the packing structure and not by any concentration gradient. One alternative approach to radial heat transfer has been to consider two- or three-dimensional ow elds, that is, to include velocity components transverse to the main direction of ow. Early attempts to obtain such ow elds were made by
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research

Article

Figure 2. Simulated packed beds for CFD analysis: (a) N = 3.96, (b) N = 5.96, (c) N = 7.99.

Table 1. Comparison of Experimental to ComputerGenerated Overall Voidage


N 3.96 5.96 7.99 (expt) 0.476 0.451 0.431 (computer model) 0.466 0.450 0.432 % deviation 2.1 0.2 0.2

where some authors replaced eq 6 with


T |r = R = Tw
(9)

Stanek and Szekely30 who substituted a two-dimensional voidage into the dierential form of the vectorized Ergun equation. Their model could not satisfy the no-slip condition on the tube wall, and also highlighted the need for a realistic voidage or other suitable measure of bed structure to obtain ow elds. This could be provided by packing network models31,32 or cell models33 or even statistical models34 which have all had computational drawbacks. Several attempts have been made to obtain two components of velocity from the twodimensional NavierStokes equations, but most of them concluded that radial velocity components were negligible in fully developed ow in the packed bed and reported results for only one-dimensional models.12,19,20 Radial heat transfer models that included two components of velocity were investigated by several authors.3538 The equations used took the general form
1 T T T r c pGz(r , z) + Gr(r , z) = k r r r r z r
kr T r = hw (T |r = R Tw )
r=R

(5)

(6)

T r

=0
r=0

(7) (8)
15246

T |z = 0 = Tin(r )

and others included axial terms. Most neglected the radial mass ux Gr(r,z), for example Stanek and Vychodil35 concluded that radial ow terms were less than 1% of the velocity magnitude, while Eigenberger36 and Froment37,38 and their co-workers stated that strong radial ow components were found only in the rst particle layer or a short entrance region. The conclusions of the previous paragraph seem to contradict the well-established idea that the main contributor to radial heat transfer, at least as Re increases, is radial displacement of uid around the particles, that is, convective dispersion. One explanation is that the uid mechanics models that were used to obtain the velocity components were all based on averaged measures of bed structure, usually voidage. Some used (r,z) and others only (r); however, all involved a degree of smoothing of the bed structure. We suggest here that the higher values of vr along the entire bed that would be needed to account for the observed rates of radial heat transfer are suppressed by the use of smoothed voidage elds in the NavierStokes equations. In fact, the void fraction at a point can have only values of zero or one, and it changes abruptly at the local level as the radial coordinate passes from particle to uid and back many times. It is these abrupt changes in void fraction that give rise to the redirection of ow, giving strong local variations in p and vr, which in turn result in the observed radial heat transfer rates. These local variations can be averaged out in smoothed or global approaches to xed bed structure and uid ow. To avoid uncertainties in the existing models of ow in xed beds and the desire to avoid premature smoothing of the velocity components at the local level, a dierent approach is to use simulations of velocity elds directly in heat transfer models. For example, Dixon et al.5 put forward a model of
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research xed-bed heat transfer that employed the uid and solid thermal conductivities of the phases and a two-dimensional velocity eld composed of the components vz(r) and vr(r,z). The velocity components were calculated from a computergenerated packing, which was transformed into a network model.31 Flows through the network branches were calculated and then averaged to obtain the velocity components. The model was able to give heat transfer through the bed center reasonably well but it was not as accurate in the wall region. One explanation was that the ow channels parallel to the wall could not be included in the network and in addition during the averaging process strong radial ows to and from the wall were added to give no net ow, although the net heat transfer was probably signicant. Problems with the network model near the tube wall prevented further development of the heat transfer model at that time. Ziolkowska and Ziolkowski, in their eective viscosity model,17 derived a ow model that was macroscopically onedimensional in the direction of the pressure drop, but which could have local radial components which were related to uid radial dispersion. In their later work18 a radial dispersion coecient was included directly into the equation of continuity and obtained by analogy from friction factor correlations. The equation of continuity was then integrated analytically to obtain the interstitial component vr. More recently Schnitzlein29 attempted to capture the uctuations in the local velocities caused by the local packing structure. His approach was to use continuum models dened in terms of a two-dimensional spatially dependent porosity. From a computer-generated sphere pack Schnitzlein obtained an angularly averaged voidage (r,z) which he used in the NavierStokes equations. The asymptotic value for the radial Peclet number was found to be more than twice as high as the experimental value of Per() = 11. Using a three-dimensional network model which did not involve averaging of the bed structure, better agreement for dispersion was found.39 Recent developments have been made in CFD which allow the simulation of ow, heat, and mass transport in full beds of spheres of several hundred particles.4043 Such computations could allow the actual local values of the velocity components vz and vr to be obtained directly by simulation, with no need for a model. Several of these studies have presented axial velocity vz(r) contours or proles, and more recent work has begun to give axial proles of area-averaged radial velocity.44 The object of the present work is to combine the model of Dixon et al.5 with CFD simulations to demonstrate the concept that kr and hw can be replaced by the velocity components vz(r) and vr(r,z), along with a model of stagnant bed conduction ke0(r), to give a more physically realistic description of xed bed radial heat transfer.

Article

to a user-specied overlap tolerance by moving spheres in turn so as to expand the bed vertically. A gravitational force is then applied downward on each sphere to compact the bed, moving the spheres so as to reduce the particle center of mass until a stopping criterion is reached, while respecting the overlap tolerance. The results have been found to usually be in closer agreement with published experimental data than previous algorithms for conned beds. We found that the random allocation of spheres in this collective rearrangement type of algorithm gave some unrealistic sphere packings at the bottom layers of the bed, especially for lower N. For these simulations, it was decided to build the packing from a base layer with spheres in a ring around the wall. We therefore combined the original algorithm with an initial position algorithm46 to more realistically locate the wall layer of spheres at the tube bottom. With some modication this solved the problem, and this combined random-deterministic algorithm was used to generate a range of sphere packs with nominal diameter dp = 0.0254 m. Three values of N were chosen for detailed study in this work, N = 3.96, 5.96, and 7.99. These values were chosen through consideration of the available experimental data, and also to cover a reasonable range of N. Side views of the three packings are included in Figure 2 to give a sense of the type of structure that can be generated, along with the relative dimensions. For the N = 3.96 bed, 250 spheres were used, the N = 5.96 bed was generated from 400 spheres and for the N = 7.99 bed there were 800 spheres in the model. For each packing, overall voidage was calculated from the nominal tube and particle diameters, the packed bed length, and the number of particles. Experimental values were obtained from the results of Mueller47 as reported in the later paper by the same author,48 and comparisons are presented in Table 1. The overall voidage in the computer-generated beds is generally lower, as would be expected from a soft-sphere algorithm.

3. DISCRETE PARTICLE (CFD) BED GENERATION CFD simulations of full beds of spheres played a major role in our methodology. We have generated a range of tube-toparticle diameter ratios (3 N 9.3) for sphere-packed beds. To do this, we adapted a published soft-sphere algorithm45 which produces sphere packs with lower voidage than the usual drop-and-roll packing algorithm. The algorithm rst places a predetermined number of spheres Np of given diameter dp at random positions inside a cylinder by allowing interpenetration between the particles. The cylinder has diameter such as to give a specied N, and a chosen initial voidage 0 sets the initial tube length. The overlaps are then reduced in the absence of gravity
15247

Figure 3. Comparison of radial bed voidage proles for N = 3.96 between experimental measurements and computer-generated sphere pack calculations.
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research The agreement is excellent for the three cases shown, with slightly higher error for lower N. Further validation of the computer-generated structures was conducted by comparison of the radial voidage prole to literature data, again from Mueller.48 The validation comparisons are shown in Figures 35, as plots of voidage as a function of

Article

Figure 6. Verication of CFD solution for supercial axial velocity vz for the N = 3.96 bed and Re = 240, using three mesh sizes.

Figure 4. Comparison of radial bed voidage proles for N = 5.96 between experimental measurements and computer-generated sphere pack calculations.

of the sphere positions for the entire bed in each case using the formulas developed by Mueller.49,50 The voidage proles for N = 3.96 are shown in Figure 3. The prole shows two minima corresponding to the two layers of spheres along the radial coordinate. Overall, the features and magnitude of the experimental () are well-reproduced. The high void fraction at the tube center ( 2) is caused by the hole down the tube center due to the packing structure. The voidage prole for N = 5.96 is shown in Figure 4; some slight shift of the prole toward the tube wall may be attributed to the soft-sphere algorithm which produces a more compacted packing. This feature was also observed in similar algorithms previously.45 The downturn for values of 3 is due to anomalies at the center of the bed, where it is dicult to dene small enough surfaces to obtain accurate values. Nevertheless, this region is very small and of lesser importance compared to the near-wall region where excellent agreement is found. The voidage prole of the N = 7.99 bed is shown in Figure 5 and shows similar features and good agreement. The magnitudes of the maxima and minima are especially accurately found by the algorithm. The low voidage at 4 (tube center) for the computer-generated packing is not an anomaly; for this particular packing the spheres lined up along the center-line. The general good agreement shown in the graphs demonstrates that the computer-generated sphere pack reproduces the essential features of experimental measurements. The locations of maxima and minima are correctly reproduced, as well as their magnitudes. This nding as well as the results for overall voidage gives us condence in our computer-generated models for the simulation of the velocity and temperatures in a xed bed.

Figure 5. Comparison of radial bed voidage proles for N = 7.99 between experimental measurements and computer-generated sphere pack calculations.

4. 3D DISCRETE PARTICLE (CFD) SIMULATION MODEL The equations for the CFD simulation of uid ow and heat transfer in a single phase in this study are the equations of conservation of mass, momentum, and energy. The conservation of mass (continuity) equation is
(ui) + = Sm t xi
15248

the dimensionless distance from the tube wall, in multiples of particle diameter. The voidage proles were calculated from a list

(10)

dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research In the simulations described here, the source term Sm was equal to zero. The equation for conservation of momentum in direction i is
ij (uiuj) (ui) p + gi + Fi = + + xj xj xi t

Article

(11)

In this equation p is the static pressure, ij is the stress tensor, and gi is the gravitational body force. For the present work the external body force term Fi was zero. The stress tensor ij for a Newtonian uid is dened by
uj ui 2 ul ij = + ij xi 3 xl xj

(12)

Here is the molecular viscosity and the second term on the right-hand side of the equation is the eect of volume dilation. The energy equation is
(uih) (h) + t xi = j hjJj u Dp T + + (ik) i + S h xi xi xi Dt xk
(13)

In this equation h is the enthalpy and for the present study the user-dened volumetric heat source term Sh was zero. Radiation was not included in the CFD simulation model due to the relatively low laboratory-level temperatures simulated. The above equations were solved in their original form for laminar ows; for turbulent ows the Reynolds-averaged Navier Stokes (RANS) models were used, with the original equations being ensemble-averaged. With RANS models all turbulence length scales are modeled. The solution variables are decomposed into mean, ui and uctuating, ui components and integrated over an interval of time that is large compared to the small-scale uctuations. When this is applied to the standard NavierStokes equations, the result is
u ui uj u ( ui ) p i + j = + + xi xj xi xj t xj 2 u ( uiu j) l + xj 3 xl

set at the tube outlet. For the energy balance, the tube wall temperature, Tw = 368.15 K and the temperature of the inlet ow Tin = 298.15 K were specied. At the particle soliduid interfaces continuity of temperature and heat ux was enforced. CFD simulations were carried out to obtain velocity and temperature elds in full beds of spheres for the three cases, N = 3.96, N = 5.96, and N = 7.99. The nominal particle diameter was 1 in. (0.0254 m) in all columns and the nominal tube diameters were 3.96 in. (0.1009 m), 5.96 in. (0.151384 m), and 7.99 in. (0.202946 m). The models had a length of 0.0254 m of empty tube before the bed inlet and a length of 0.0508 m of empty tube after the bed to be able to place the inlet and outlet boundary conditions away from the packing. The packed lengths of the columns were as given in Figure 2. Simulations were run over a range of ow rates to give Re in the range 802000. The uid for the CFD simulations was taken as air with constant properties corresponding to a bed average temperature of 333.15 K. These were density = 1.059545 kg/m3, viscosity = 2.0291 105 kg/ms, specic heat cp = 1800 J/kgK, and thermal conductivity kf = 0.0287 W/mK. The particles were taken to be alumina with properties as density s = 1947 kg/m3, specic heat cps = 1000 J/kgK, and thermal conductivity ks = 1.0 W/mK. The model geometries and the mesh were constructed using the commercial software GAMBIT 2.4.6, with the help of journal les to carry out the repetitive creation and placement of the spheres. To obtain a ne enough near-wall mesh for the k- method we used boundary layer prism cells at outside particle surfaces and at the tube walls; tetrahedral cells were used in the main uid volume and inside the particles. The unstructured tetrahedral mesh cell size was 1.524 103 m (dp/16.7) and the boundary layer mesh thickness was 2.54 105 m (dp/1000) with three layers on the tube wall and a single layer on the particle surfaces. The N = 3.96, 5.96, and

(14)

The velocities and other solution variables are now represented by Reynolds-averaged values, and the eects of turbulence are represented by the Reynolds stresses, ( uiu j ). To close the system of equations the Reynolds stresses are put in terms of the averaged ow quantities. In the present work we used a k- two-equation model, which is a two-zone model designed to be integrated all the way to the wall, provided that a suciently ne mesh is used there. Descriptions of the k- turbulence model are available in standard references and will not be repeated here. Laminar ow models were used for the three lowest ow rates simulated, and a turbulent ow model was used at the highest ow rate, which corresponded to Re = 1900. Boundary conditions for the momentum di erential equations were provided by taking the no-slip condition on all solid surfaces, both tube wall and particles. A uniform velocity prole was used at the tube inlet, and a pressure of 1 atm was
15249

Figure 7. Verication of nite element solution of 2D pseudocontinuum model with constant coecients against analytical solution at dierent bed depths.
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research

Article

Figure 8. Contours of (a) axial velocity (m/s), (b) radial velocity (m/s), and (c) temperature (K), in the x = 0 plane of the N = 3.96 xed bed and for Re = 240. Dotted boxes indicate regions used for close-up velocity vectors shown in Figure 9.

Figure 9. Close-up analysis of boxed regions from Figure 8 with N = 3.96 and Re = 240: (a) velocity vectors colored by axial velocity (m/s), (b) velocity vectors colored by radial velocity (m/s), and (c) temperature contours (K).

7.99 total mesh sizes were 8.904 million, 15.579 million, and 28.47 million cells, respectively. To remove the problem of meshing around the contact points between the particles and between the tube wall and the particles, the technique of shrinking the diameters of the particles to 99% of the original diameter was used, so that the particles had an actual diameter 0.025146 m. To provide the same size gaps for the particlewall contact points the tube diameters were all increased by 2.54 104 m also. This decision implied that heat transfer by particleparticle or particle-wall area contacts was not represented in this model. Other approaches to the problem of meshing around contact points have been developed, and these were recently reviewed and compared.51 Although the use of gaps between particles does aect heat transfer uxes and temperature distributions, in this study the same simplication was made in applying the formula for the eective stagnant thermal
15250

conductivity in the pseudocontinuum vzvr model, so that the comparisons were made on the same basis. To verify mesh independence, a mesh renement study was carried out on the velocity proles in the N = 3.96 tube for Re = 240. Three mesh sizes were compared, the base case size of 1.524 103 m (dp/16.7), and two ner meshes of 1.27 103 m (dp/20) and 1.016 103 m (dp/25). The three corresponding proles of supercial velocity vz(r) are presented in Figure 6 where they are shown to coincide almost exactly except for a small region at the bed center, where the velocity is higher due to the hole in the packing along the centerline which is typical for N = 4 beds. This shows that the base case mesh (dp/16.7) is acceptable for the present study of velocity elds. The governing equations described above were solved using the nite volume commercial CFD code FLUENT 6.3.26. The pressure-based segregated solver was used, with the SIMPLE
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research

Article

Figure 10. Cross-section of N = 7.99 xed bed column showing radial surfaces and angular planes used for averaging and sampling of the CFD results.

Figure 11. Averages over angular surfaces in the N = 3.96 bed, Re = 240: (a) void fraction, (b) temperature (K), (c) radial velocity (m/s), (d) axial velocity (m/s).

scheme for pressure-velocity coupling. First-order upwind interpolation was used for the convection terms; all diusion terms used second-order discretization. Tests using secondorder upwind interpolation for N = 3.96 and Re = 240 showed that the velocity prole was unchanged for the most part, except for small dierences very close to the tube wall and also
15251

in the bed center. This result implies that rst-order upwind interpolation is adequate for calculating velocity proles, but will likely not be sucient for calculations of pressure drop. Under-relaxation factors were left at the FLUENT default settings, unless some instability was observed in the iterations, when they were occasionally reduced. The convergence was
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research monitored by the temperature values at the bed exit, and checking the energy balance for the column, in addition to the residuals. Simulations were run on a Sun Microsystems X2200 M2 x64, a 64-bit server with two dual-core processors (4 CPU total) at 2.6 GHz each with 8 GB RAM. Several experiments have been carried out in our group in order to validate CFD simulations by comparison to experiments for xed bed heat transfer. During an experiment under typical laboratory conditions, a heated tube wall packed bed was used52 with N = 2 at relatively low ow rates of air corresponding to Re < 2000. Thermocouples were used to acquire temperature proles as functions of radial position. CFD was used to model the ow and heat transfer. A very good quantitative as well as an excellent qualitative agreement between CFD simulation and experimental results was obtained for heat transfer of the N = 2 xed bed column in the laboratory, after allowance was made for unavoidable dierences between the experimental and simulation situations. In the later study,53 CFD simulations of heat transfer in xed beds of spheres at higher ow rates were validated by comparison to experimental measurements in a pilot-scale rig. The comparisons were made for particle Reynolds numbers in the range 2200 < Re < 27000 for a tube-to-particle diameter ratio of N = 5.45, and for particle Reynolds numbers in the range 1600 < Re < 5600 for N = 7.44. CFD models of a 0.20 m heated packed length, consisting of 304 spheres for N = 5.45 and 722 spheres for N = 7.44, were solved, corresponding to the experimental setup. The CFD simulations compared well to the experimental data: trends with Re, N, and bed depth were captured, and the quantitative agreement of temperature proles was reasonable, allowing for the diculties of obtaining experimental data in larger equipment at industrially applicable ow rates. The results of these two validation studies give condence for the present work that CFD simulations can yield realistic temperature elds in wall-heated packed beds.

Article

accounting for the Smoluchowski eect, as none of these phenomena were included in the CFD model. It was felt necessary to validate the use of the Zehner Schlunder formula on a pointwise basis. Although others have used this formula in this way, we are not aware of any rigorous demonstration of its validity. The formula was developed and tested only on bed average data by the original authors, with some empiricism introduced which would not necessarily hold locally. Our recent work54 on stagnant radial conduction in an annular xed bed conrmed the pointwise use of the Zehner Schlunder formula for the conditions of the present study. For the convective contribution, we postulated that the dispersion of heat by radial displacement of ow could be simulated by a two-dimensional velocity eld using the components vz(r) and vr(r,z). The new 2D model equations are (for a pseudohomogeneous case):
T T 1 T 0 rk (r ) + vr(r , z) = c pvz(r ) e z r r r r
T |r = R = Tw
(17) (18)

T r

=0
r=0

(19) (20)

T |z = 0 = Tin(r )

5. NEW 2D VELOCITY-BASED PSEUDOCONTINUUM (vzvr) MODEL In the new modeling approach the near-wall decrease in thermal resistance due to changes in conduction and radial mixing is not lumped at the wall, but instead is linked directly to the conduction and uid ow phenomena causing the changes. The conduction contribution to the heat transfer model was obtained with stagnant thermal conductivity, ke0(r) from the ZehnerSchlunder cell model,6 as a function of true uid thermal conductivity, particle thermal conductivity, and the radial bed voidage prole. The formula was used in the form
ke0(r ) = k f 1 1 + 1 2 NM

COMSOL nite element multiphysics software was used to solve this pseudocontinuum two-dimensional pseudohomogeneous heat transfer model. The coecients were implemented by the incorporation of tables for void fraction and axial velocity as functions of r, and radial velocity as a function of r and z, which were then interpolated as input to the nite element method. Standard model verication tests were run by mesh renement and demonstrated mesh independence. To verify the nite element model, eqs 1720 were adapted to be the same as eqs 14, solved, and compared with the analytical series solution of those equations. Conditions were based on laboratory experimental data for 0.5-in. ceramic spheres in a 2-in. diameter column (N = 4) with mass ow rate Gz = 1.982 kg/m2s, with parameters Bi (hwR/kr) = 1.2075 and kr/kf = 103.654. The inlet feed was air at 298.15 K with the wall temperature at 368.15 K. The analytical solution and nite element model temperatures at three dierent bed depths (z/L = 1/3, z/L = 2/3, and z/L = 1) were compared, as shown in Figure 7. The symbols represent the analytical solution temperatures; the curves represent the nite element computed proles. The nite element solution and the analytical solution showed excellent agreement for heat transfer in a low-N xed bed with the same boundary conditions and heat transfer coecient. The computed radial temperature proles coincided almost exactly.

kp 1 kp B B+1 B1 ln 2 kp B N M 2 (N M ) (15)

where kp = ks/kf, N M = 1 B/kp and

1 10/9 B = 1.25

(16)

In this formulation the terms for radiation and heat transfer through nite area contacts were omitted, as well as the term
15252

6. AVERAGING AND SAMPLING THE 3D CFD MODEL The CFD simulations give the 3D distributions of void fraction (r, , z), velocity components vz(r, , z), and vr(r, , z) and temperature T(r, , z), the latter in both uid and solid. A major task in this research was to analyze the full 3D velocity eld to obtain a manageable representation in terms of velocity components, by data sampling and averaging. It is widely accepted that the important variation in temperature is given by T(r, z). Note that the usual techniques of volume averaging in xed beds will not apply here as there is no separation of length scales.
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research 6.1. Temperature and Velocity Contour Analysis on a Single Plane. The axial and radial velocity and temperature contours on a single plane at constant in the z- and r-directions for the N = 3.96 xed bed column and Re = 240 are presented in Figure 8 panels a, b, and c, respectively. In these gures, only the central portion of the bed is presented, to avoid anomalies due to end eects. This corresponded to 0.1397 z 0.4191 m, approximately ve particle layers from the inlet and three particle layers from the bed exit.

Article

Figure 12. Supercial axial velocity proles for the three dierent values of N, Re = 240.

The axial velocity in Figure 8a is high at the center of the bed where the bed voidage is close to unity and also near the tube wall at about r/R = 0.98. It then decreased to zero because of the wall boundary layer no-slip condition. In addition, the axial velocity plot shows that the ow velocities near the particle surfaces are zero, as dened by the no-slip condition on all the solid surfaces in the geometry. Comparisons of the axial velocity vz to the total velocity magnitude showed that the axial velocity was usually the dominant component. When there was a dierence between vz and |v|, it meant that another component of ow played a signicant role, and in this case it was the radial velocity vr. The radial velocity in Figure 8b exhibited positive and negative velocities extending over a range from 0.75 to 0.82 m/s. Although most of the values clustered around zero, many small regions could be seen with positive and negative velocities up to 0.3 m/s in magnitude. The positive radial velocities mean those velocity vectors which moved from center of the bed to the tube wall, and negative radial velocities mean those velocities which moved from tube wall to the center of the bed. Radial velocity had a signicant eect on the temperature distribution at a local level in the bed; when radial velocity was negative and the ow direction was to the center of the bed then high temperature uid owed from heated tube wall and penetrated into the center of bed. The temperature contours in Figure 8c show that the radial temperature prole did not develop smoothly from the inlet to the outlet of the bed. The temperature contours had development and reduction in the radial position. This was due to the eect of the radial velocity that dominated the radial

Figure 13. Average of radial velocity contours in xed bed columns of (a) N = 3.96, (b) N = 5.96, and (c) N = 7.99 for Re = 80.
15253
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research heat transfer mechanism, supplemented by the uidparticle uid conduction path, rather than to the axial velocity. The dotted-line box superimposed on the temperature contours in Figure 8c highlights the uneven propagation of the temperature into the center of the bed. Corresponding dotted boxes were also marked in Figure 8a and Figure 8b for the velocity plots. Velocity vector plots represent the uid velocity magnitude and direction at each control volume. To study the local velocities corresponding to the dotted line box in Figure 8c, the velocity vectors are presented in Figure 9a colored by axial velocity and in Figure 9b colored by radial velocity. Figure 9c shows the temperature development and reduction more clearly in the line box. The velocity vectors colored by axial velocity were high at the center and close to the tube wall in the low voidage area where the distance between particles or between particles and the wall was larger. The velocity vectors colored by radial velocity had negative and positive values depending on the particle distributions. The radial velocity vectors illustrate how the temperature distribution to the center of the bed was changed locally in the radial position. When the radial velocities were negative and the velocity vectors left the tube wall to the center of the bed (between particles 4, 5 and 6), the temperature proles were more developed in the center of the bed due to the enhanced transfer of high temperature uid from the heated tube wall. However when the radial velocities were positive and the velocity vector approached the tube wall the temperature proles were reduced (between particles 1, 2, and 3) as the heat transfer from the tube wall was inhibited by the local motion of the uid.

Article

The analysis on this single plane emphasizes the importance of the local radial velocity components, and the necessity for sampling and averaging methods to avoid cancellation eects which would diminish the eect of the local velocity uctuations. 6.2. Coarse-Graining the 3D Velocity Fields. The void fraction and axial velocity vz were extracted from averaging of dierent cylindrical planes in the radial direction inside the xed bed, as illustrated in Figure 10 at the left. These two quantities were therefore averaged over both angular () and axial (z) coordinates. In the z-direction, again only the center parts of the beds were used, to avoid end eects, and these corresponded to 0.1397 z 0.4191 m for the N = 3.96 bed, 0.127 z 0.296 m for the N = 5.96 bed, and 0.0762 z 0.3048 m for the N = 7.99 bed. To determine radial velocity as a function of radial and axial position, the straightforward extension of the averaging method to radial cylindrical planes subdivided into small increments in the axial direction was not used. Instead, we dened 32 angular planes from the center of the bed to the tube wall at 11.25 spacing. Figure 10 on the right shows the position of the angular planes inside the xed bed. All velocities were averaged at the same r and z position in all the angular planes together. But the cells in the dierent angular planes were not located at the same r and z positions because of the dierent unstructured tetrahedral meshes in each plane. Therefore we used interpolation for all planes to have values of the radial velocities at the same r and z positions and then averaged them. In this case 200 (radial) 200 (axial) points were extracted for each plane, and then velocity components at the corresponding

Figure 14. Average of radial velocity contours in xed bed columns of (a) N = 3.96, (b) N = 5.96, and (c) N = 7.99 for Re = 240.
15254
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research points of all planes were averaged. The number of angular planes used varied depending on the N value and on Re, as it was found that fewer planes could sometimes be used with good results. This averaging procedure gave better resolution of vr(r,z) than other methods that were tried, which was especially important to capture the variation of the radial velocity near the tube wall. 6.3. Local Features in the Averaged Fields. Figure 11 shows the angular averages of bed voidage, temperature, radial velocity, and axial velocity for N = 3.96 at Re = 240, in the (r,z) plane. The gures show the entire tube length, including the unpacked inlet and outlet sections. As can be seen in Figure 11a the 2D voidage distribution in the bed represented conrmation that assuming the voidage prole as a function of only radial direction was a reasonable assumption, since the high bed voidage area was located at the center of the bed and close to the tube wall for the entire axial direction. The axial velocity prole followed the voidage prole and it was not changed signicantly in the axial direction so that it was also safely assumed to be a function of radial position only (Figure 11d). The radial velocity, in contrast in Figure 11c had a variety of dierent values in both axial and radial directions, conrming that it had to be assumed as a function of both r and z directions. The averaging damped out the more extreme local values shown in Figure 8b for a single plane, but retained sucient variation to represent the radial heat transfer, as illustrated in the next section. The two-dimensional averaged temperature distribution is also shown for comparison, in Figure 11b. The region of lower temperature at the center of

Article

the outlet area corresponds to an area of recirculating uid behind the packing.

7. RESULTS OF COMPARISON OF MODELS For the comparisons of vz() shown in Figure 12, only the middle parts of the packed beds were used; that is, the cylindrical surfaces were clipped to the z-values listed in section 6.2. For all the two-dimensional comparisons of vr(r,z) and T(r,z) shown in Figures 1320 again only the middle parts of the packed beds were used, and the z-coordinate was reset to zero at the start of the sample region. 7.1. Axial Velocity Proles. The radially varying axial velocity vz(r) was extracted from the averaging of dierent cylindrical surfaces in the radial direction inside the xed bed. The averaged axial velocity corresponding to Re = 240 is compared in Figure 12 for all three values of N. The axial velocity followed closely the voidage prole, therefore slow velocities were located in high void fraction regions but in the wall vicinity the axial velocity increased and then decreased due to the boundary layer and the no-slip condition at the wall. Overall, when plotted in terms of , the three proles of vz are very similar, with maxima and minima in the same locations and of comparable magnitude (note that the supercial velocity was the same in each case). Some dierences may be seen between the N = 3.96 prole and the other two proles in the rst two velocity peaks from the wall, possibly due to the special structure at the lowest N. For each value of N, there is some anomalous behavior at the bed center caused by the high or low void fractions there as discussed previously. In this plot the values of within 0.1 from the bed center were discarded

Figure 15. Average of radial velocity contours in xed bed columns of (a) N = 3.96, (b) N = 5.96, and (c) N = 7.99 for Re = 950.
15255
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research

Article

Figure 16. Average of radial velocity contours in xed bed columns of (a) N = 3.96, (b) N = 5.96, and (c) N = 7.99 for Re = 1900.

due to the diculty in obtaining meaningful averages with very small radial surfaces. 7.2. Radial Velocity Fields. Radial velocities vr(r,z) were obtained from the averaging of dierent angular planes at dierent angular positions. The average radial velocity contours for the N = 3.96, 5.96, and 7.99 xed bed columns are presented in Figure 13, panels a, b, and c, respectively, for Re = 80, and in the corresponding three parts of Figure 14 for Re = 240, again for N = 3.96, 5.96, and 7.99. Similarly the radial velocity elds for the three N at the transitional value Re = 950 are given in Figure 15 and under turbulent ow conditions at Re = 1900 in Figure 16. Considering each Re value separately, there does not appear to be a great deal of similarity between the three N values for each ow rate. Both N = 5.96 and N = 7.99 show higher values of vr near the bed center, and more extreme values than does N = 3.96. This may be attributed to the stronger axial bypassing both along the tube wall and down the bed center which is peculiar to the N = 3.96 bed structure. If the four dierent vr(r,z) elds corresponding to the four dierent Re are compared for each N individually, some patterns start to emerge. The regions at high positive vr for N = 5.96 and N = 7.99 are in the same positions near the centerline. They alternate axially with patches of strongly negative vr. High magnitudes of vr appear to occur in bands in between the layers of particles arranged against the tube wall. Regions of vr close to zero correspond to averages at positions which are mostly inside the layers of particles. It is possible to discern two, three, and four axial bands corresponding to the two, three, and four particles along the bed radius for N = 3.96, 5.96, and 7.99, respectively.
15256

It would be expected that vr would alternate in sign in both the axial and radial directions as the overall averages must come to zero, there being no net radial ow on the scale of the bed radius. It appears that more analysis and more values of N and Re will be needed to develop either empirical or mechanistic approaches to predicting vr. For the purposes of the present work, tables of values from the averaged CFD simulations are sucient. 7.3. Radial and Axial Temperature Fields. The comparisons between the CFD 3D discrete particle and 2D pseudocontinuum vzvr model temperatures are presented in Figure 17 for Re = 80. The CFD discrete particle model temperature contours in Figure 17 were obtained from averaging of the same 32 angular planes that were used for the radial velocity. As can be seen, the pseudocontinuum model temperature contours had excellent quantitative agreement with the averaged CFD temperature contours for all three N values. The model predicted the axial temperature distribution fairly well; in addition the radial temperature distribution was predicted very well. The temperature of the xed bed had a rougher distribution in the axial direction due to the particle heat transfer by conduction and the radial velocities distribution between particles. Both models showed development of the temperature into the center of the bed at the same locations. At this lowest of the ow rates there is signicant thermal penetration into the bed, across the entire radius for N = 3.96 and 5.96, and across most of the radius for N = 7.99. The temperature elds for N = 3.96 have a slight dierence at the bed center where the higher void fraction and axial velocity give more rapid temperature development in the CFD simulation.
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research

Article

Figure 17. Comparison of CFD 3D discrete particle model temperature contours and 2D vzvr pseudocontinuum model temperature contours in xed bed column of (a) N = 3.96, (b) N = 5.96, and (c) N = 7.99 for Re = 80.

Figure 18. Comparison of CFD 3D discrete particle model temperature contours and 2D vzvr pseudocontinuum model temperature contours in xed bed column of (a) N = 3.96, (b) N = 5.96, and (c) N = 7.99 for Re = 240.
15257
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research

Article

Figure 19. Comparison of CFD 3D discrete particle model temperature contours and 2D vzvr pseudocontinuum model temperature contours in xed bed column of (a) N = 3.96, (b) N = 5.96, and (c) N = 7.99 for Re = 950.

Figure 20. Comparison of CFD 3D discrete particle model temperature contours and 2D vzvr pseudocontinuum model temperature contours in xed bed column of (a) N = 3.96, (b) N = 5.96, and (c) N = 7.99 for Re = 1900.
15258
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research For N = 5.96 the two temperature elds are very similar, with the CFD temperature level slightly lower at the outlet, but otherwise the comparison is good. For N = 7.99 the two temperature elds are in excellent agreement, with even smaller details the same. This demonstrates that the new model can predict the convection and conduction heat transfer locally with the pseudocontinuum energy equation very well in a 2D model for low Reynolds numbers for which conduction and convection both play signicant roles. The temperature contours for Re = 240 for all N are shown in Figure 18. It is seen that at this higher ow rate there is less penetration of the thermal front into the bed than for the lower ow rate. The temperature contours between CFD and the vzvr model were in generally good agreement. However, close to the center of the bed, the CFD temperature developed more than in the vzvr model; this was especially so for N = 3.96 and somewhat the case for N = 5.96. Since the center of the bed was taken as a symmetric boundary condition in the vzvr model the velocity prole close to the center had to satisfy this boundary condition limitation and could not develop in the same way as for the CFD model. For N = 7.99 the two temperature elds were again in excellent agreement. The temperature contours at the much higher ow rate Re = 950 for all three N are shown in Figure 19. The simulations for the near-turbulent regime Re = 950 could be run as either laminar or turbulent models in the CFD; in this work Re = 950 was assumed as laminar ow and the ow from CFD showed reasonable results. The results are more sensitive to the averaging method for higher Reynolds number compared to the low Reynolds numbers since the radial velocities have lower negative and higher positive values, so it was necessary to average more angular surfaces to avoid low or high radial velocities at dierent points, which would then cancel with the averaging method. Averaging of more angular surfaces extracted better radial velocities to be used in the pseudocontinuum vzvr model. This may account for the stronger appearance of discrete temperature features in the N = 3.96 vzvr model results, although generally the near-wall comparisons were good. For N = 5.96 the vzvr temperature is a little low, and also for N = 7.99. The extent of the temperature contours and the general shape of the developing contours are both good in all three cases. The temperature contours for turbulent ow at Re = 1900 for all N are shown in Figure 20. The CFD and pseudocontinuum model temperature comparison showed some slight dierences in the results. This was due to the dimensional reduction of 3D to 2D. There is a wider wall region of high temperature for the vzvr model than in the CFD simulations, but comparisons show very good results for all three N for the temperature levels across the tube radius, in terms of both extent and magnitude. Overall, the 2D pseudocontinuum heat transfer model based on the velocity elds from CFD produced very reasonable results compared to the 3D CFD temperature simulations without the need to introduce any adjustable parameters such as kr/kf and hw or an eective viscosity.

Article

medium cell model applied pointwise to account for heat transfer by thermal conduction and its dependence on local bed voidage. Fluid ow elds in validated xed beds of spheres of tube to particle ratio N = 3.96, 5.96, and 7.99 were obtained by solving the 3D NavierStokes equations in a detailed CFD approach which preserved the actual bed structure in the simulation. A methodology was developed to obtain the axial velocity, vz(r), and radial velocity, vr(r,z) from the 3D discrete particle results. Stagnant eective thermal conductivity was calculated at each radial position from the Zehner-Schlunder model as function of local bed porosity, uid thermal conductivity, and solid thermal conductivity. Comparisons were made for Reynolds numbers in the range 801900, for the three values of N, under typical laboratoryscale conditions that would be used with a steam-heated column. The temperatures calculated by the new 2D velocitybased heat transfer equation showed very good quantitative and qualitative agreement with the values given by the detailed CFD simulation. The trends with Re and N were captured well. The results of this study suggest that the local radial velocity components can account for the convective contributions to radial heat transfer in a packed bed of spheres. They are not negligible if computed from CFD simulations in model geometries that preserve the discrete bed structure instead of replacing it with a pseudocontinuum or eective medium. As the ultimate objective is a computationally tractable 2D pseudocontinuum reaction engineering model, care needs to be taken in averaging the information from the 3D discrete simulations for use in lower-dimensional eective models. Earlier approaches that began from eective medium models with smoothed measures of bed structure substituted into the volume-averaged NavierStokes equations or their equivalent to obtain velocity elds, all concluded that radial velocity components were negligible as the smoothed structure led to the result that (p/r) 0 and thus to vr 0, as the local radial variations in pressure and velocity were averaged out. The axial and radial velocity proles obtained in this study suggested that it may be possible to obtain generalized velocity components for use in a predictive model.

AUTHOR INFORMATION

Corresponding Author Present Address

*E-mail: agdixon@wpi.edu. Department of Chemical Engineering, Massachusetts Institute of Technology (MIT), Cambridge, MA 02139, USA.
Notes

The authors declare no competing nancial interest.

8. CONCLUSIONS The main object of this work was to demonstrate the feasibility of modeling radial temperature proles in xed beds of spheres without using any adjustable parameters such as kr/kf and hw, and without using eective heat conduction approaches for uid mechanical phenomena. Instead, radial heat transfer was to be predicted using local position-dependent components of axial and radial velocity to represent heat transfer by uid motion and its decrease near the tube wall, and a local eective
15259

ACKNOWLEDGMENTS This material is based upon work supported by the National Science Foundation under Grant No. CTS-0625693. NOMENCLATURE cp = uid specic heat, J/(kgK) cps = solid specic heat, J/(kgK) B = shape parameter for ZehnerSchlunder formula Bi = wall Biot number, hwR/kr dp = particle diameter, m dt = tube diameter, m Fi = external body force per unit volume, kg/(m2s)
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research gi = body force in i-direction per unit mass, m/s2 Gr = radial mass ux, kg/(m2s) Gz = axial mass ux, kg/(m2s) h = specic enthalpy, J/kg hi = specic enthalpy of species i, J/kg hw = apparent wall heat transfer coecient, W/(m2K) Ji = mass diusive ux of species i, kg/(m2s) k = turbulent kinetic energy, J/kg ke0 = stagnant eective thermal conductivity, W/(mK) kf = uid thermal conductivity, W/(mK) kp = thermal conductivity ratio, ks/kf kr = eective radial thermal conductivity, W/(mK) ks = solid thermal conductivity, W/(mK) L = bed length, m N = tube-to-particle diameter ratio, dt/dp NM = parameter for ZehnerSchlunder formula Np = number of particles in computer-generated packing p = static pressure, Pa Per() = limiting value of Peclet number (Gzcpdp/kr) at high Re r = radial coordinate, m R = tube radius, m Re = Reynolds number based on particle diameter, dpvz/ Sh = energy source term, J/(m3s) Sm = mass source term, kg/(m3s) t = time, s T = temperature, K Tin = inlet temperature, K Tw = wall temperature, K ui = generic velocity component in direction i, m/s u0 = supercial plug-ow velocity, m/s vr = radial velocity component, m/s vz = axial velocity component, m/s xi = coordinate direction i, m z = axial coordinate, m
Greek Letters

Article

(1) Leal, L. G. Challenges and opportunities in uid mechanics and transport phenomena. In Perspectives in Chemical Engineering Research and Education, Advanced Chemical Engineering; Colton, C.K., Ed.; Academic Press: Boston, MA, 1991, 16, 6179. (2) Dixon, A. G. Fixed bed catalytic reactor modelingThe radial heat transfer problem. Can. J. Chem. Eng. 2012, 90, 507527. (3) Dixon, A. G.; Nijemeisland, M. CFD as a design tool for fixedbed reactors. Ind. Eng. Chem. Res. 2001, 40, 52465254. (4) Dixon, A. G.; Nijemeisland, M.; Stitt, E. H. Packed tubular reactor modeling and catalyst design using computational fluid dynamics. Adv. Chem. Eng. 2006, 31, 307389. (5) Dixon, A. G.; Ng, K. M.; Chu, C.-F. Heat transfer in packed tubes with a small tube-to-particle diameter ratio. AIChE Annual Meeting, Nov. 610, 1989, San Francisco, CA.
15260

= bed voidage 0 = initial voidage for bed generation = eective thermal conductivity of the uid (molecular and turbulent), W/(mK) = angular coordinate, radians = uid viscosity, kg/(ms) = dimensionless distance from tube wall, (Rr)/dp = uid density, kg/m3 s = solid density, kg/m3 ij = viscous ux of j-momentum in the i-direction, kg/ms2 = specic dissipation rate, s1

REFERENCES

(6) Zehner, P.; Schlunder, E.-U. War meleitfah igkeit von Schuttungen bei mas sigen Temperaturen. Chem. Ing. Tech. 1970, 42, 933941. (7) Hatta, S.; Maeda, S. Heat transfer in beds of granular catalystI. Chem. Eng. (Jpn.) 1948, 12, 5658. (8) Coberly, C. A.; Marshall, W. R., Jr. Temperature gradients in gas streams flowing through fixed granular beds. Chem. Eng. Prog. 1951, 47, 141150. (9) Ziolkowska, I.; Ziolkowski, D. Fluid flow inside packed beds. Chem. Eng. Process. 1988, 23, 137164. (10) Cheng, Z.-M.; Yuan, W.-K. Estimating radial velocity of fixed beds with low tube-to-particle diameter ratios. AIChE J. 1997, 43, 13191324. (11) Vortmeyer, D.; Schuster, J. Evaluation of steady flow profiles in rectangular and circular packed beds by a variational method. Chem. Eng. Sci. 1983, 38, 16911699. (12) Bey, O.; Eigenberger, G. Fluid flow through catalyst filled tubes. Chem. Eng. Sci. 1997, 52, 13651376. (13) Vafai, K. Convective flow and heat transfer in variable-porosity media. J. Fluid Mech. 1984, 147, 233259. (14) Tien, C. L.; Hunt, M. L. Boundary-layer flow and heat transfer in porous beds. Chem. Eng. Process. 1987, 21, 5363. (15) Whitaker, S. The Forchheimer equation: A theoretical development. Transp. Porous Media 1996, 25, 2761. (16) Subagyo; Standish, N.; Brooks, G. A. A new model of velocity distribution of a single-phase fluid flowing in packed beds. Chem. Eng. Sci. 1998, 53, 13751385. (17) Ziolkowska, I.; Ziolkowski, D. Modeling of gas interstitial velocity radial distribution over a cross-section of a tube packed with a granular catalyst bed. Chem. Eng. Sci. 1993, 48, 32833292. (18) Ziolkowska, I.; Ziolkowski, D. Modelling of gas interstitial velocity radial distribution over cross-section of a tube packed with granular catalyst bed; effects of granule shape and of lateral gas mixing. Chem. Eng. Sci. 2007, 62, 24912502. (19) Eisfeld, B.; Schnitzlein, K. A new pseudo-continuous model for the fluid flow in packed beds. Chem. Eng. Sci. 2005, 60, 41054117. (20) Guo, B.; Yu, A.; Wright, B.; Zulli, P. Simulation of turbulent flow in a packed bed. Chem. Eng. Technol. 2006, 29, 596603. (21) Winterberg, M.; Tsotsas, E.; Krischke, A.; Vortmeyer, D. A simple and coherent set of coefficients for modelling of heat and mass transport with and without chemical reaction in tubes filled with spheres. Chem. Eng. Sci. 2000, 55, 967979. (22) Winterberg, M.; Tsotsas, E. Correlations for effective heat transport coefficients in beds packed with cylindrical particles. Chem. Eng. Sci. 2000, 55, 59375943. (23) Sundaresan, S.; Amundson, N. R.; Aris, R. Observations of fixedbed dispersion models: The role of the interstitial fluid. AIChE J. 1980, 26, 529536. (24) Hinduja, M. J.; Sundaresan, S.; Jackson, R. A crossflow model of dispersion in packed bed reactors. AIChE J. 1980, 26, 274281. (25) Klingman, K. J.; Lee, H. H. Alternating flow model for mass and heat dispersion in packed beds. AIChE J. 1987, 33, 366381. (26) Stewart, W. E. Transport phenomena in fixed-bed reactors. Chem. Eng. Prog. Symp. Ser. 1965, 61, 6165. (27) Kronberg, A. E.; Westerterp, K. R. Nonequilibrium effects in fixed-bed interstitial fluid dispersion. Chem. Eng. Sci. 1999, 54, 3977 3993. (28) Iordanidis, A. A.; van Sint Annaland, M.; Kronberg, A. E.; Kuipers, J. A. M. A critical comparison between the wave model and the standard dispersion model. Chem. Eng. Sci. 2003, 58, 27852795. (29) Schnitzlein, K. Modelling radial dispersion in terms of the local structure of packed beds. Chem. Eng. Sci. 2001, 56, 579585. (30) Stanek, V.; Szekely, J. Three-dimensional flow of fluids through nonuniform packed beds. AIChE J. 1974, 20, 974980. (31) Chu, C. F.; Ng, K. M. Flow in packed tubes with a small tube to particle diameter ratio. AIChE J. 1989, 35, 148158. (32) Thompson, K. E.; Fogler, H. S. Modeling flow in disordered packed beds from pore-scale fluid mechanics. AIChE J. 1997, 43, 13771389.
dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Industrial & Engineering Chemistry Research


(33) Jiang, Y.; Khadilkar, M. R.; Al-Dahhan, M. H.; Dudukovic, M. P. Single phase flow modeling in packed beds: Discrete cell approach revisited. Chem. Eng. Sci. 2000, 55, 18291844. (34) Ziolkowska, I.; Ziolkowski, D. Stochastic approach to modelling and analysing of gas flow field in tubes packed randomly with spherical catalyst pellets. Chem. Eng. Process 2005, 44, 11671180. (35) Stanek, V.; Vychodil, P. Mathematical model and assessment of thermally induced gas flow inhomogeneities in fixed beds. Chem. Eng. Process. 1987, 22, 107115. (36) Daszkowski, T.; Eigenberger, G. A reevaluation of fluid flow, heat transfer, and chemical reaction in catalyst filled tubes. Chem. Eng. Sci. 1992, 47, 22452250. (37) Delmas, H.; Froment, G. F. A simulation model accounting for structural radial non-uniformities in fixed bed reactors. Chem. Eng. Sci. 1988, 43, 22812287. (38) Papageorgiou, J. N.; Froment, G. F. Simulation models accounting for radial voidage profiles in fixed bed reactors. Chem. Eng. Sci. 1995, 50, 30433056. (39) Schnitzlein, K. Modeling radial dispersion in terms of the local structure of packed beds. Part II: Discrete modeling approach. Chem. Eng. Sci. 2007, 62, 49444947. (40) Bai, H.; Theuerkauf, J.; Gillis, P.; Witt, P. A coupled DEM and CFD simulation of flow field and pressure drop in fixed bed reactor with randomly packed catalyst particles. Ind. Eng. Chem. Res. 2009, 48, 40604074. (41) Magnico, P. Pore-scale simulations of unsteady flow and heat transfer in tubular fixed beds. AIChE J. 2009, 55, 849867. (42) Baker, M. J.; Tabor, G. R. Computational analysis of transitional air flow through packed columns of spheres using the finite volume technique. Comput. Chem. Eng. 2010, 34, 878885. (43) Eppinger, T.; Seidler, K.; Kraume, M. DEM-CFD simulations of fixed bed reactors with small tube to particle diameter ratios. Chem. Eng. J. 2011, 166, 324331. (44) Zobel, N.; Eppinger, T.; Behrendt, F.; Kraume, M. Influence of the wall structure on the void fraction distribution in packed beds. Chem. Eng. Sci. 2012, 71, 212219. (45) Salvat, W. I.; Mariani, N. J.; Barreto, G. F.; Martinez, O. M. An algorithm to simulate packing structure in cylindrical containers. Catal. Today 2005, 107108, 513519. (46) Mueller, G. E. Numerical simulation of packed beds with monosized spheres in cylindrical containers. Powder Technol. 1997, 92, 179183. (47) Mueller, G. E. Radial void fraction distributions in randomly packed fixed beds of uniformly sized spheres in cylindrical containers. Powder Technol. 1992, 72, 269275. (48) Mueller, G. E. Numerically packing spheres in cylinders. Powder Technol. 2005, 159, 105110. (49) Mueller, G. E. Radial porosity in packed beds of spheres. Powder Technol. 2010, 203, 626633. (50) Mueller, G. E. A simple method for determining sphere packed bed radial porosity. Powder Technol. 2012, 229, 9096. (51) Dixon, A. G.; Nijemeisland, M.; Stitt, E. H. Systematic mesh development for 3D CFD simulation of fixed beds: Contact points study. Comput. Chem. Eng. 2013, 48, 135153. (52) Nijemeisland, M.; Dixon, A. G. Comparison of CFD simulations to experiment for convective heat transfer in a gassolid fixed bed. Chem. Eng. J. 2001, 82, 231246. (53) Dixon, A. G.; Walls, G.; Stanness, H.; Nijemeisland, M.; Stitt, E. H. Experimental validation of high Reynolds number CFD simulations of heat transfer in a pilot-scale fixed-bed tube. Chem. Eng. J. 2012, 200202, 344356. (54) Dixon, A. G.; Gurnon, A. K.; Nijemeisland, M.; Stitt, E. H. CFD testing of the pointwise use of the ZehnerSchlunder formulas for fixed-bed stagnant thermal conductivity. Int. Commun. Heat Mass Transfer 2013, 42, 14.

Article

15261

dx.doi.org/10.1021/ie4000568 | Ind. Eng. Chem. Res. 2013, 52, 1524415261

Potrebbero piacerti anche