Sunteți pe pagina 1din 13

Chemical Engineering Science 61 (2006) 4341 4353 www.elsevier.

com/locate/ces

CFD study on particle-to-uid heat transfer in xed bed reactors: Convective heat transfer at low and high pressure
A. Guardo a , , M. Coussirat b , F. Recasens a , M.A. Larrayoz a , X. Escaler b
a Chemical Engineering Department, ETS dEnginyeria Industrial de Barcelona, Universitat Politcnica de Catalunya, Av. Diagonal 647,

ETSEIB 08028, Barcelona, Spain


b Fluid Mechanics Department, ETS dEnginyeria Industrial de Barcelona, Universitat Politcnica de Catalunya, Av. Diagonal 647,

ETSEIB 08028, Barcelona, Spain Received 26 September 2005; received in revised form 30 January 2006; accepted 1 February 2006 Available online 23 February 2006

Abstract Computational uid dynamics (CFD) has proven to be a reliable tool for xed bed reactor design, through the resolution of 3D transport equations for mass, momentum and energy balances. Solution of these equations allow to obtain velocity and temperature proles within the reactor. The numerical results obtained allow estimating useful parameters applicable to equipment design. Particle-to-uid heat transfer coefcient is of primal importance when analyzing the performance of a xed bed reactor. To gain insight in this subject, numerical results using a modied commercial CFD solver are presented and particle-to-uid heat transfer in xed beds is analyzed. Two different congurations are studied: forced convection at low pressure (with air as circulating uid) and mixed (i.e., free + forced) convection at high pressure (with supercritical CO2 as circulating uid). In order to impose supercritical uid properties to the model, modications into the CFD code were introduced by means of user dened functions (UDF) and user dened equations (UDE). The obtained numerical data is compared to previously published data and a novel CFD-based correlation (for free, forced and mixed convection at high pressure) is presented. 2006 Elsevier Ltd. All rights reserved.
Keywords: Computational uid dynamics; Packed bed; Fluid mechanics; Heat transfer; Hydrodynamics; Supercritical uids

1. Introduction The design of a xed bed reactor requires an extensive knowledge of heat transfer characteristics within the bed. As an example, within a xed bed reactor the evaluation of temperature prole as well as the heat transfer rates is essential in the control and performance of the reactor. In general, uid temperature is measured with little difculty, but the measurement of the solid surface temperature is not easy. The particle temperature or the temperature drop at the particle surface then has to be estimated in terms of the heat transfer coefcient between the particle and the uid. Because of the importance of the particle-to-uid heat transfer coefcient in xed bed reactors, a considerable effort has been made to evaluate this parameter. Experimental
Corresponding author. Tel.: +34 934016676; fax: +34 934017150.

E-mail address: alfredo.guardo-zabaleta@upc.edu (A. Guardo). 0009-2509/$ - see front matter 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.ces.2006.02.011

determination of heat transfer coefcient for a wide variety of systems have been made using various experimental techniques, under either steady-state or unsteady-state conditions. Some theoretical studies have also been carried out to explain the observed experimental results. There is, of course, no exact theory which describes in a suitable way the transport phenomena in a xed bed. An extensive review of the aforementioned can be found in Wakao and Kaguei (1982). Although there is broad information available of published experimental data for particle-to-uid forced convection heat transfer at low pressure (see e.g. Wakao and Kaguei, 1982), no data are available for high pressure situations or supercritical xed bed reactors. Many authors have studied catalytic reactions using a supercritical uid as a solvent (Poliakoff et al., 1996; Ramirez et al., 2004), but main interest has been centered in kinetics of chemical reactions under supercritical conditions. Other authors have studied mass transfer in xed beds under supercritical conditions, but applied to supercritical extraction (Debenedetti and Reid, 1986; Stber et al., 1996). The recent

4342

A. Guardo et al. / Chemical Engineering Science 61 (2006) 4341 4353

study and development of new compact packed bed supercritical nuclear reactors opens a branch for the study of heat transfer phenomena under high pressure conditions (see e.g. Oka et al., 1994). Modeling and simulation are essential tools in design and scale-up of xed bed reactors, and as the performance requirements of this equipments is growing, it is needed that the proposed model would be able to dene not only spatial distribution of involved species, but also temperature and velocity proles within the reactor. A main limitation of heat transfer models in xed beds is the poor knowledge about ow pattern within them. Experimental studies have been carried out to look at ow maldistribution in packed columns, nding that distribution of uid inside the column is far from being uniform, showing channeling of ow in the column wall (Yin et al., 2000). Dispersion and ow in porous media have been studied with magnetic resonance imaging (MRI) techniques. This technique has allowed identifying experimentally stagnant ow regions, radial ows and back-ow regions within the xed bed (Manz et al., 1999; Suekane et al., 2003). This conrms the idea of the presence of a strong radial ow at the uid inlet into the bed, developing preferential ow zones where heat transfer would be accentuated, creating temperature gradients along the bed. Computational uid dynamics (CFD) has proven to be a reliable tool when modeling and simulating ow and heat transfer phenomena for designing process equipment. Nowadays, CFD can be applied to the more physical aspects of chemical engineering, cases in which the knowledge of heat and mass transfer around complexes geometries is essential to a good understanding of the involved phenomena. Several reviews on the applicability of CFD methodology to chemical reactors and process equipment design have been published in the last years (Bode, 1994; Harris et al., 1996; Kuipers and Van Swaaij, 1998; Ranade, 2002; Joshi and Ranade, 2003). Early CFD work on discrete particle simulations of xed beds included 2D studies that resolved ow patterns and heat transfer around proposed ideal geometries (Dalman et al., 1986; Lloyd and Boehm, 1994), but proposed geometries in these analyses are too far from reality to be considered as representative of a xed bed. Other studies have used variational methods for evaluating steady-state velocity proles for isothermal, incompressible ow in 2D packed beds, showing that there is channeling in the wall when the particle-to-channel diameter ratio is low (Vortmeyer and Schuster, 1983). Evolution on computers design and performance has allowed researchers to build up more complex three-dimensional models, overcoming the difculty associated to the grid generation and the computational power required to solve such models. Wall-to-uid heat transfer in sphere packed beds has been studied around different geometries (Logtenberg et al., 1999; Dixon and Nijemeisland, 2001). These studies focused on using CFD to obtain heat transfer parameters such as the wall Nusselt number, and gave reasonable agreement with experimental estimates. Magnico (2003) applied an Eulerian and a Lagrangian approach to model the hydrodynamics and transport properties of packed beds. Inuence of the turbulence model on heat transfer prediction, near-wall treatment and numerical studies

of the laminar, transition and turbulent ow zones in a packed bed have also been analyzed through wall-to uid heat transfer models, showing the importance of the denition of a good mesh around the heat transfer surfaces in order to properly capture the associated boundary layer phenomena (Guardo et al., 2004, 2005). Nijemeisland and Dixon (2004) have presented a detailed study of the relationship between the local ow eld and the local wall heat ux in a packed bed of spheres, nding that local heat transfer rates do not correlate statistically with the local ow eld. Pressure drop and particle-to-uid heat and mass transfer in low channel to particle diameter ratio have also been studied (Calis et al., 2001; Romkes et al., 2003), showing that a commercial CFD code can be used to adequately predict particle-touid heat transfer of a single free sphere, and that the CFD code can predict the heat/mass transfer characteristics for packed beds of spherical particles with a channel to particle diameter ratio between 1 and 2 with an average error of 15% compared to experimental values. Recently, Gunjal et al. (2005) have presented a computational study of a single phase ow in packed beds of spheres, analyzing the inuence of the packing array on ow mixing and heat transfer parameters. The main purpose of this work is to review CFD capabilities for predicting particle-to-uid heat transfer coefcients when a supercritical uid is used as a solvent in a xed bed reactor. First of all, numerical simulations are presented for a validation model case (single sphere model), and secondly, convective particle-to-uid heat transfer at low pressure is presented. Results obtained are used to analyze mesh dependence of the numerical results at low ow velocities. Finally, mixed convection at high pressure is modeled and analyzed. Numerical results obtained are compared to broadly accepted correlations and a novel CFD-based correlation for particle-to-uid heat transfer at high pressure is presented. 2. Mesh design and CFD modeling In order to properly understand the transport mechanisms present in the studied cases, a dimensionless analysis under work conditions of the set of used equations has been developed. Dimensionless equations corresponding to mass, momentum and energy balances are as follows:
0

t0 St

j jt ju jt

u0 L

) = 0 (u

u0 ), ( u L

(1)

)u + (u

= Eu( p) +

1 + ( u) { [ u T ]} Re 1 1 + ( u) + { [ u T ]} + ( fb ), ReT Fr = Ec )T + (u + 1 1 1 )] + [v( : u Re Re Pr 1 1 Re PrT T k T . k T

(2)

St

jT jt

(3)

A. Guardo et al. / Chemical Engineering Science 61 (2006) 4341 4353

4343

In order to gain insight on the problem, the order of magnitude of the dimensionless groups was estimated taking physicalchemical properties values for the circulating uids from experimental data and empirical correlations available in the literature (Reid et al., 1987; Yaws, 1999; Poling et al., 2000). Boundary (operating) conditions for each analyzed situation are shown in Table 1. Turbulence intensity used to estimate ReT was set in 1%, and Reynolds analogy was used to estimate PrT from ReT (White, 1991). For rst analysis purposes, PrT was assumed as a constant value within the bed, justied in the fact that majority of experimental results shows a range of variations between 0.75 < PrT < 2 for air and water (Kays, 1994). Results of the order of magnitude of the dimensionless groups are shown in Table 2. In the case of particle-to-uid heat transfer at low pressure, dimensionless analysis allows identifying the problem as forced convection heat transfer in laminar or turbulent ow. For Eq. (2) it becomes clear that viscous forces decrease their contribution as Re increases. Inertial gravity forces increase their contribution as Re decreases, and pressure drop together with turbulent forces become the most important terms in Eq. (2) at high Re. In Eq. (3), the convective and the diffusive term become important for energy balance. According to this, the mesh should be designed to properly dene a minimum cell size to capture in a suitable way the computed variables on the proposed geometry. When using Reynolds-averaged NavierStokes (RANS) equations, grid renement leads to a solution closer to the exact solution of a system of equations that implies considerable physical approximations (Shur et al., 2005). According to this, a mesh sensitivity analysis should be done. For both balances, steady-state analysis can be used.
Table 1 Boundary (operating) conditions for analyzed cases Boundary condition Circulating uid Temperature at the inlet, K Temperature at packing surface, K Pressure, Pa Mass ow at the inlet, kg/m2 s Velocity at the inlet, m/s Low pressure Air 298 423 101 325 3 104 7.5 101 High pressure CO2 330 340 1 107 0.0130.132

When analyzing a heat transfer case at high pressure, the problem is identied as mixed (free + forced) convection in laminar or transitional ow. Body forces and pressure drop become the most important terms in Eq. (2). Turbulence forces contribution to Eqs. (2) and (3) is negligible. For these cases, mesh should properly dene the boundary layer around the proposed geometry. In order to couple energy and momentum balance, unsteady state analysis is required. 3. Geometrical models In order to validate heat transfer calculations, a single sphere suspended in a box was created. With this numerical modeling it was tried to t CFD data to a generally accepted theoretical model (Ranz and Marshall, 1952) that assumes one sphere in an innite uid. In order to keep the model reasonable in size, in the CFD model the innite uid was limited in a box with a square ow inlet plane of seven sphere diameters and a length of 16 sphere diameters (Nijemeisland, 2000). Models with ow inlet planes with sizes of 2, 3, 4, 6, 8 and 9 diameters were also created in order to discard the presence of wall effects on temperature and velocity proles. An unstructured tetrahedral mesh is built in the uid region. No mesh is built in the sphere interior. The sphere in the box was designed with the same dimensions as the spheres used in the xed bed model, detailed below. For the heat transfer simulations at low and high pressure, a xed bed model was created. A 44 spheres stacking with a sphere-to-tube diameters ratio of 3.923 was chosen to be the geometrical model (Guardo et al., 2004). Modeled geometry has been constructed following the bottom-up technique (generating surfaces and volumes from nodes and edges) in order to control mesh size around critical points (i.e., particleto-particle and particle-to-wall contact points), necessary to avoid grid elements skewness, and also to gain computational resources by reducing the number of elements in zones of low interest (i.e., geometrical zones away from contact points). The construction of wall-to-particle and particle-to-particle contact points is also an important subject in model generation. Previous work reports the emulation of contact points (leaving small gaps between surfaces and assuming zero velocity in the gap) to avoid convergence problems (Nijemeisland, 2000). In this study, to include real contact points, the spheres were modeled

Table 2 Dimensionless groups magnitude orders Low pressure Re St Eu Fr Pr Br Ec ReT PrT 100 102 103 104 104 101 1010 1010 101 101 101 101 101 103 104 102 101 1010 108 1010 108 100 101 102 100 101 101 102 100 101 108 106 108 106 100 101 103 101 101 100 101 106 106 100 101 High pressure 100 105 106 1010 100 1016 1016 101 101 101 104 104 108 100 1014 1014 100 101

4344

A. Guardo et al. / Chemical Engineering Science 61 (2006) 4341 4353

Fig. 1. Detail of the contact points in the xed bed geometrical model.

overlapping by 0.5% of their diameters with the adjacent surfaces in the geometric model (see Fig. 1). Convergence problems were not detected during simulation runs. 4. Models analysis NavierStokes equations together with energy balance were solved using commercially available nite volume code software Fluent 6.2. For the validation model and the low pressure simulations, the uid was taken to be incompressible, Newtonian, and in a laminar or turbulent ow regime. Air at standard conditions was chosen as the simulation uid. Incompressible ideal gas law for density and power law for viscosity were applied to the model for making these variables temperaturedependent. For the high pressure simulations, the uid was taken to be Newtonian, in laminar ow regime and with variable density. CO2 was chosen in this case as the simulation uid, and its properties at high pressure were incorporated to the solver code through user dened functions (UDF) and user dened equations (UDE). PengRobinson equation of state was used to calculate uids density and heat capacity, Lucas method was used to calculate the viscosity, and thermal excess method was used to calculate the thermal conductivity (Poling et al., 2000). Simulations were run in an HP C3000 Workstation, and simulation times ranged from 1 to 240 h depending on the studied case. Numerical convergence of the model was checked based on a suitable diminution of the normalized numerical residuals of all computed variables. For a more complete convergence checking the average static temperature at the bed outlet were also chosen as monitors in all cases. 5. Results and discussion 5.1. Validation model: one sphere suspended in an innite uid The rst step in the solution of xed bed ow and heat transfer problems in a complex geometry was to solve the problem

of one sphere suspended in an innite domain of uid. The main goals of this exercise were to check the mesh sensitivity and to compare numerical solution with the theoretical solution proposed by Ranz and Marshall (1952) in order to check the feasibility for obtaining suitable numerical results for this problem. Particle surface temperature was set at 400 K and the uid inlet temperature and the box wall temperature was set in 300 K. Under-relaxation factors for pressure, momentum and energy were initially set to 0.05, 0.1 and 0.2, respectively (Gunjal et al., 2005), and increased progressively after convergence until values of 0.2, 0.4 and 0.8, respectively. In the case of turbulent ow simulations, under-relaxation factors for turbulent quantities were set in 0.4. A rst order discretization scheme for pressure, momentum and energy equations was used until convergence was achieved, and the results obtained were used as initial solution for a new simulation applying a second order discretization scheme for momentum and energy equations. A test case was developed in order to discard the presence of wall effects in temperature and velocity proles over particle surface. For a single velocity condition (Re 300), velocity and temperature proles near the particle surface were studied in a box with a square ow inlet plane of 2, 4, 6, 7, 8 and 9 sphere diameters. Fig. 2 shows the velocity prole and Fig. 3 the temperature prole in the central plane of the sphere. As it can be seen, wall effects over velocity prole (Fig. 2) are present until a four sphere diameters inlet plane is used. In the case of temperature proles, wall effects were only detected when a two sphere diameters inlet plane is used (Fig. 3). According to this, the selection of seven sphere diameters inlet plane is justied and no disturbing wall effects on heat transfer are expected. For each simulation, temperature contour plots (Fig. 4) were analyzed, and heat ux through the particle surface was recorded. With this data the heat transfer coefcient (h) could be determined: q = h Ae (Tp T ). (4)

From the values of h, the Nusselt number (Nu) was computed and compared with the theoretical solution proposed by Ranz and Marshall (1952). Concerning to the mesh sensitivity analysis the test performed consisted in changing the mesh density in the particle surface in order to properly capture the boundary layer associated problem. Four simulation sets were obtained using different cell sizes at the particle surface. Inlet velocity was varied for each simulation set (0.33 < Re < 3300), and Nu was obtained for the mentioned range. Fig. 5 shows the numerical results obtained. As it can be seen, low mesh density at particle surface can lead to erroneous solutions due to an incorrect denition of the boundary layer. As mesh density increases, tting with theoretical solution improves; results obtained for the two ner meshes are almost identical, so it can be established that simulations have reached an asymptotic solution. Optimal mesh densities at the particle surface were recorded and used to build the mesh for the xed bed model. Agreement between numerical and theoretical results was considered to be satisfactory.

A. Guardo et al. / Chemical Engineering Science 61 (2006) 4341 4353


10

4345

Wall Distance [in particle diameters]


1 0.001

0.01

0.1

Velocity [m/s]

2D

4D

6D

7D

8D

9D

Fig. 2. Wall effects test. Velocity prole over particle surface at Re 300.
10

Wall Distance [in particle diameters]


1 100

1000

Temperature [K]
2D 4D 6D 7D 8D 9D

Fig. 3. Wall effects test. Temperature prole over particle surface at Re 300.

5.2. Fixed bed model: forced convection heat transfer at low pressure The main purpose of this set of simulations was to test CFD capabilities in a particle-to-uid heat transfer xed bed model. Standard correlations for Nu and experimental data were selected as reference values to compare against the numerical results generated (Wakao and Kaguei, 1982). For each simulation, inlet velocity was varied (0.2 < Re < 900) and Nu was obtained. Within this range of Re, laminar, transition and turbulent ow regime were expected (Jolls and Hanratty, 1966). For the modeling in the turbulent ow region (Re > 300), previous experience shows that the SpalartAllmaras turbulence model is a good choice (Guardo et al., 2005).

The two ner mesh densities at the particle surface used in the validation model were used to build the xed bed model. Although prior experience with the validation model indicates that the selected meshes should be enough to capture heat transfer phenomena over the particle surface, and that the Redependence of the mesh is weak for RANS modeling, some inuence of the mesh size can be expected when the Re increases (Spalart, 2000). Anticipating this situation, two meshes (one coarser and one ner than the aforementioned) were also prepared and tested in order to assure mesh-independence of the obtained results. Boundary conditions for the modeling are described in Table 1. Under-relaxation factors were set as described for the validation case. A rst order discretization scheme for pressure, momentum and energy equations was used until

4346

A. Guardo et al. / Chemical Engineering Science 61 (2006) 4341 4353

convergence was achieved, and the results obtained were used as initial solution for a new simulation applying a second order discretization scheme for momentum and energy equations.

In packed beds with a constant particle surface temperature, the resistance to heat transfer resides only on the uid side. Theoretical analysis shows that if axial uid thermal dispersion is taken into consideration, then, the heat balance equation becomes u d Tf 3h (1 ) (Tf Tp ) = + Cp dx
ax

d 2 Tf . dx 2

(5)

With the following boundary conditions: u(Tf T0 ) = dTf =0 dx


ax

d Tf dx

at x = 0 (inlet),

(5a) (5b)

at x = xL (outlet).

Fig. 4. Temperature elds for the single sphere model. (A) Re = 0.33; (B) Re = 3396. Fluid ows in the positive z-axis direction.

For each simulation, velocity, temperature and density proles along the bed were recorded and analyzed. With the collected data, h was obtained from Eq. (5). The values of the axial uid thermal dispersion coefcient ( ax ) used in Eq. (5) are estimated from Wakao (1976). From the obtained values of h, Nu was computed and compared with the correlation proposed by Wakao et al. (1979). Results are shown in Fig. 6. It can be noticed that in the laminar and transition ow zone (Re < 300), the results do not show dependence on the mesh density. At lower Reynolds numbers (Re < 10), results shows that the tting against Wakaos correlation is not good. For a single velocity condition different meshes give results in a wide range of Nu and no relation with mesh density can be established in any case. Experimentally, one-shot measurements have demonstrated that no denite Nu values can be obtained at low Re (Wakao, 1976). The fact that any Nu value within the obtained ranges yields approximately the same value of ax suggests that at this low Re, particle-touid heat transfer makes little contribution to the overall heat

40

30

Nu

20

10

0 0.1 1 10 100 1000 10000

Re

V cell / V p = 7.81 x 10-4 V cell / V p = 3.20 x 10-5

-4 V cell / Velements 60040 p = 5.75 x 10 Ranz & Marshall

V = 2.09 x 10-4 143725 cell / Vp elements

Fig. 5. Nup vs. Re for the single sphere model.

A. Guardo et al. / Chemical Engineering Science 61 (2006) 4341 4353


100

4347

Nu

10

1 0.1 1 10 Re 100 1000

V cell / V p = 5.77 x 10-4 V cell / V p = 3.04 x 10-5

V = 2.09 x 10-4 104325 cell / V p elements Wakao et al.

V = 3.21 x 10-5 348065 cell / V p elements


RANGE OF EXPERIMENTAL DATA (Reviewed by Wakao et al.)

Fig. 6. Nu vs. Re for the low pressure convective heat transfer xed bed model.

transfer in the system, which analytically has been reected in an increased condence range for the selected correlation at low Re (Shent et al., 1981). Using a laminar model, a good accuracy for Nu values was obtained within 10 < Re < 100, which reinforces the idea that a similar methodology can be applied to model mixed convection heat transfer at high pressure in the same Re range. For higher values of Re there is a divergence between the results obtained for tested meshes, probably due to the fact that at higher Re, turbulent transport term in the transport equation becomes more important. In RANS modeling, the balance equations possess a smooth exact solution, and the numerical solution approaches that solution as we rene the grid. The aim of grid renement is numerical accuracy (Spalart, 2000; Shur et al., 2005). Therefore, for our specic study case, an accurate turbulence modeling requires a denser mesh around the particles surface in order to capture in a more suitable way the involved turbulence phenomena in the boundary layer (Guardo et al., 2005). A divergence in the results obtained with the low density meshes and the high density meshes can be seen for Re > 300 in Fig. 6. The results obtained with the ner meshes t better the prediction of Wakao et al. (1979) in the turbulent ow zone (Re > 300), probably due to a better capture of the vorticity energetic scales associated effects. 5.3. Fixed bed model: mixed (free + forced) convection heat transfer at high pressure A numerical modeling of a xed bed where mixed convection phenomena exist was also made. The same geometrical model as the aforementioned case is used but now the circulating uid and the values of pressure and velocity are modied. The effects of density, ow direction and velocity on heat transfer are presented. Obtained numerical results were tted,

and a novel correlation based on the convective mass transfer correlation presented by Stber et al. (1996) is proposed. Two simulation sets (upow and downow operation) were done, and for each simulation, inlet velocity was varied and Nu was obtained for 9 < Re < 100. A laminar ow model was used as viscous model for all simulations. Gravitational acceleration was added to the operating conditions of the model. Under-relaxation factors were set as previously explained for the aforementioned models. A rst order discretization scheme for momentum and energy equations and a second order discretization scheme for pressure was used until convergence was achieved, and the results obtained were used as initial solution for a new simulation applying a second order discretization scheme for momentum and energy equations. Heat balance equation was used to obtain the values of Nu for each studied case. Boundary conditions for the model are described in Table 1. The presence of high density gradients lead to the formation of hydrodynamical instabilities (i.e., stream differentiation, countercurrent ow, recirculation) caused by buoyancy effects (Benneker et al., 1998). In order to compute correctly the ow elds and the heat transfer phenomena in these zones, the optimal grid used for the prior model was locally rened for each case studied around the zones where high density gradients were found. Dimensionless analysis indicated that for the studied velocity range, laminar or transition ow could be expected. Churchill (1983) suggests that in a mixed convection heat transfer situation the transition between laminar and turbulent ow can be found at 1 109 < GrH < 1 1010 . In order to corroborate this idea, the numerical results obtained for Nu for the different situations analyzed were plotted in a ow map according to Churchills criteria (Fig. 7). It can be noticed that the results obtained follow the expected behavior according to the dimensionless analysis, and transition between laminar and turbulent

4348

A. Guardo et al. / Chemical Engineering Science 61 (2006) 4341 4353

20
Laminar Flow Transition Turbulent Flow

15

Nu

10

0 1.00E+07

1.00E+08

1.00E+09

1.00E+10

1.00E+11

GrH
Assisting Flow Opposing Flow

Fig. 7. Flow map according to Churchills (1983) criteria.

ow shows up in the results by means of a high increase in Nu for alike values of GrH . 5.3.1. Effect of density gradients and ow stability Density gradients control the uid movement at low velocities. Since the early experiments of Hill (1952), there has been a growth in the literature published on hydrodynamical instabilities due to density and viscosity variations in porous media (Homsy, 1987; Manickam and Homsy, 1995). Benneker et al. (1998) proved that the hydrodynamical instabilities caused by buoyancy effects in a xed bed when density increases with height lead to a signicant increase in the ow axial dispersion within the bed. While simulating the high pressure data set, the presence of hydrodynamical instabilities captured by the model caused numerical instability in the results. The presence of large density gradients may result in convergence problems when solving the momentum, continuity and pressure equations (Lakshminarayana, 1991). As mentioned before, in order to minimize the effects of ow instability on calculations, the mesh was locally rened in each case trying to avoid the presence of high density and velocity gradients within a single cell. Numerical instability increased as inlet mass ow imposed as boundary condition decreased. Fig. 8 compares velocity elds (colored by density) for the high pressure and low pressure heat transfer situations aforementioned. It can be seen that for the high pressure situation (Fig. 8A) the presence of high density gradients (differences up to 40 kg/m3 ) lead to the differentiation of two streams: heavy (cold) uid sinking into the bed, and light (hot) uid oating out of the bed. This fact will be important when analyzing the effects of ow direction in free convection heat transfer. When analyzing the velocitydensity prole for the low pressure situation (Fig. 8B), it becomes clear that despite the fact a small density gradient is present in the results, ow eld is controlled by the inlet velocity.

Fig. 8. Velocity eld (colored by density) in an axial cut. (A) CO2 at 1107 Pa and Re = 41; (B) Air at 101 325 Pa and Re = 531.

A. Guardo et al. / Chemical Engineering Science 61 (2006) 4341 4353

4349

5.3.2. Effect of ow direction As mentioned before, ow instability zones within the bed appear if a density gradient is present along the bed. The direction of this density gradient and the direction of the ow have strong effects on the axial dispersion and the mixing of the ow (Benneker et al., 1998). Either if the ow direction assists the transfer phenomena (when density decreases with height) or is opposed to it (when density increases with height), it can be noticed that, under certain conditions, ow direction have a direct effect on heat and mass transfer phenomena. Several authors have reported in experiments that under supercritical conditions, ow direction affects extraction rates increasing them when ow direction assists the mass transfer (e.g. Sovov et al., 1994; Stber et al., 1996; Saiz et al., 2000; Germain et al., 2005). When gravity acceleration is added to the heat transfer model, the heat homologue of the aforementioned phenomenology is captured by CFD analysis. Results obtained (see Fig. 9) reect that heat transfer coefcient values are signicantly greater in assisting ow cases than those obtained for opposing ow cases. The heat transfer rising in assisting ow is favored by the temperature gradient direction (the same as ow direction), which increases ow mixing and heat transfer rates. The presence of an adverse density gradient (when temperature gradient direction is opposed to ow direction) diminishes heat transfer rates and lower coefcients are obtained. 5.3.3. Effect of ow velocity Flow velocity has a directly proportional effect over heat transfer when forced convection takes place. An increase on ow velocity leads to an increase on kinetic energy. This fact generates a better mixing and a greater heat transfer coefcient (see Fig. 6). When analyzing a mixed convection case, the obtained results indicate that despite the fact that the forced convection component of the phenomena follows this behavior,
20

free convection component is almost constant for all cases (see Fig. 10). Is clear that free convection component is intrinsically related to density and temperature elds and it is not affected by velocity. Therefore, the dominant dimensionless group describing this phenomenon will be GrH rather than Re. Independently of ow direction, for the covered range of Re heat transfer coefcient value increases almost linearly with velocity (see Fig. 9). This increase in the overall heat transfer is due exclusively to the increase of the contribution of forced convection to the total heat transfer rate. The aforementioned idea can be noticed explicitly in Figs. 11 and 12. Fig. 11 shows the contribution of natural convection to the total Nusselt number as a function of the Reynolds number. As expected, low Re numbers correspond to transport dominated by natural convection, but for Re around 100, still 30% of transport is due to natural convection (for both assisting and opposing ow). Fig. 12 shows a reduction of the numerical data points to a natural convection-free basis only depending on the Re numbers. While the contribution of free convection to total heat transfer rate decreases with Re (as its value remains almost a constant), forced convection increases, leading to greater heat transfer coefcients. 5.3.4. Correlating the obtained numerical results The obtained numerical results validates the idea that a modied correlation based on the one proposed by Stber et al. (1996) for mixed convection mass transfer under supercritical conditions in xed beds can be used to describe heat transfer phenomena in a xed bed under mixed convection regime at high pressures. Such correlation can be useful when designing supercritical xed bed reactors, particularly in the design of direct-cycle supercritical-water-cooled fast nuclear reactors (Oka et al., 1994), nowadays under study and development process.

16

12

Nu
8 4 0 40 50 60 70 80 90 100

Re
Assisting flow Opposing flow

Fig. 9. Nu vs. Re for assisting and opposing ow simulations.

4350

A. Guardo et al. / Chemical Engineering Science 61 (2006) 4341 4353

100

Nu

10

1 1 10
Re
Free Convection Forced Convection

100

Fig. 10. Nu vs. Re for free and forced convection in assisting ow.
100

Nu free / Nu (%)
10 1 10 100

Re Assisting Flow Opposing Flow

Fig. 11. Contribution of natural convection as a function of Reynolds number.


10.00

[(Nu - Nu0 ) (Nufree - Nu0 )]/Pr 0.3

1.00

0.10 1 10 100

Re
Assisting Flow Opposing Flow

Fig. 12. Contribution of forced convection to total heat transfer rate for all the runs.

A. Guardo et al. / Chemical Engineering Science 61 (2006) 4341 4353

4351

The parity plot for predicted vs. obtained Nusselt numbers is given in Fig. 13. In summary, the correlations recommended are Nu Nu0 = |NuForced (NuFree Nu0 )|. (6)

obtained with original mass transfer correlation, that showed an AARD = 18.9% (see Fig. 14). 6. Conclusions CFD proves to be a reliable tool when modeling convective heat transfer phenomena in xed beds. It allows to analyze either free or forced convection situations and the obtained results can be compared qualitatively and quantitatively against previous published data. A simple model (a sphere suspended in a box) was used as validation tool to test the capabilities of the solver reproducing an analytical solution. Preliminary tests Response sensitivity to mesh density over particle surface was studied and an asymptotic answer was obtained. Numerical results obtained were compared against the theoretical answer for estimating the heat transfer coefcient obtained by Ranz and Marshall (1952), obtaining a good agreement between numerical and theoretical answers. Forced convection at low pressure in a xed bed was simulated, and the inuence of velocity over heat transfer could be analyzed. For values of Re > 10 it was obtained a good agreement with the correlation presented by Wakao et al. (1979). At lower Reynolds numbers (Re < 10), results shows that the tting against correlation is not good. For a single velocity condition different meshes give results in a wide range of Nu and no relation with mesh density can be established in any case. No mesh sensitivity was noticed for the laminar and transition ow zones, but in the turbulent ow zone a good denition of the mesh around the particles surface is of primal importance in order to capture the turbulence vorticity energetic scales associated effects. Mixed (free + forced) convection at high pressure in a xed bed was also analyzed. For a supercritical uid in laminar ow regime it was possible to study the effects of the density gradient, ow direction and velocity over heat transfer. It was

Taking Nu0 = 2 (Wakao and Kaguei, 1982), the correlations for free and forced convection are NuFree = Nu0 + 0.001(Gr Pr)0.33 Pr0.244 , NuForced = 0.269Re0.88 Pr0.3 , (6a) (6b)

valid for the following ranges of dimensionless numbers: 9 < Re < 96, 2.2 < Pr < 3.3, and 1 108 < GrH Pr < 4 1010 . An advantage of this correlation is that a single set of parameters predicts heat transfer for assisting ow free convection (with the minus sign in Eq. (6)) as well as opposing ow free convection (with plus sign in Eq. (6)). Predicted values for Nusselt number show that estimation error is within that
20

15

Nu from CFD

10

0 0 5 10 15 20

Nu from correlations
Assisting Flow Opposing Flow

Fig. 13. Comparison of predicted vs. CFD heat-transfer data using the nal correlation.

25

20

15

Nu T
10 5 0 1 2 3 4 5 6 7 8 9 10 11

Simulation run
CFD Correlation (Eq. 7)

Fig. 14. Estimated error for correlation against simulation runs.

4352

A. Guardo et al. / Chemical Engineering Science 61 (2006) 4341 4353

noticed that the presence of large density gradients conditioned the mesh inuence over the numerical results when computing the mixing and the heat transfer within the computed domain. An adverse density gradient generates hydrodynamical instabilities that produce an increase of the axial dispersion and a diminished heat transfer rate. This fact caused numerical instability in the simulation process. In order to eliminate numerical instability, the mesh was locally rened in each case trying to avoid the presence of high density and velocity gradients along the bed. Inuence of ow direction over heat transfer was also analyzed and it was noticed that in assisting ow regime (upow operation), greater heat transfer rates are obtained. The rising in the density along the bed height due to the adverse density gradient helps the axial dispersion to grow, obtaining less mixing within the bed and smaller heat transfer coefcients. Flow velocity also affects the heat transfer rate. The value of Nu increases almost linearly with ow velocity and this increase is due exclusively to the increase in the contribution of forced convection to the overall heat transfer. Free convection is independent of ow velocity and its value remains almost constant within the studied velocity range. The obtained numerical results validates the idea that a modied correlation presented in this work, based on the one proposed by Stber et al. (1996) for mixed convection mass transfer under supercritical conditions in xed beds, can be used to describe heat transfer phenomena in a xed bed under mixed convection regime at high pressures. A novel correlation (Eq. (6)) is presented for estimating the free and forced convection effects and Nu from Re, Pr and GrH . Presented correlations are valid for 9 < Re < 96, 2.2 < Pr < 3.3, and 1 108 < GrH Pr < 4 1010 . Notation Ae Br Cp dp Ec Eu fb Fr g Gf GrH h k kT L Nu NuForced NuFree P Pr PrT effective area of heat transfer, m2 Brinkman number ( u2 /kf T ) specic heat, J kg1 K 1 particle diameter, m Eckert number (Br/Pr) Euler number ( P / u2 ) body forces, kg m3 Froude number (u2 /L g) gravity forces, m s2 mass velocity, kg m2 s1 Grashof heat number (g T L3 2 / 2 ) heat transfer coefcient, W m2 K 1 thermal conductivity, W m1 K 1 Eddy thermal conductivity, W m1 K 1 characteristic length, m Nusselt number (h dp /kf ) forced convection Nusselt number free convection Nusselt number pressure, Pa Prandtl number (Cp /kf ) Prandtl turbulent number (Cp /kT )

q RT Re ReT St t T Tf Tp T u Vcell Vp x

heat ux, W m2 tube radius, m Reynolds number (dp u / ) Reynolds turbulent number (dp u / Strouhal number (L/t0 u) time, s temperature, K uid temperature, K particle surface temperature, K room temperature, K velocity, m s1 cell volume, m3 particle volume, m3 axial distance variable, m

T)

Greek letters
ax

axial uid thermal dispersion coefcient, m2 s1 bed void fraction dynamic viscosity, kg m1 s1 Eddy viscosity, kg m1 s1 kinematic viscosity, m2 s1 uid density, kg m3 stress tensor, N m2

Sub/superindex 0 characteristic quantity dimensionless quantity

Acknowledgments A fellowship to A. Guardo from the FI program (DURSIGeneralitat de Catalunya, Spain/European Social Fund) is acknowledged. Funding from the Spanish Ministry of Science and Technology (Grant no. AGL2003-05861) is also appreciated. References
Benneker, A., Kronberg, A., Westerterp, K., 1998. Inuence of buoyancy forces on the ow of gases through packed beds at elevated pressures. A.I.Ch.E. Journal 44, 263270. Bode, J., 1994. Applications of computational uid dynamics in the chemical industry. Chemical Engineering & Technology 17, 145148. Calis, H.P.A., Nijenhuis, J., Paikert, B.C., Dautzenberg, F.M., van den Bleek, C.M., 2001. CFD modeling and experimental validation of pressure drop and ow prole in a novel structured catalytic reactor packing. Chemical Engineering Science 56, 17131720. Churchill, S., 1983. Single phase convective heat transfer. In: Schlnder, E. et al. (Eds.), Heat exchanger design handbook. Hemisphere publishing corporation, New York, NY, USA. Dalman, M.T., Merkin, J.H., McGreavy, C., 1986. Fluid ow and heat transfer past two spheres in a cylindrical tube. Computers & Fluids 14, 267281. Debenedetti, P., Reid, R.C., 1986. Diffusion and mass transfer in supercritical uids. A.I.Ch.E. Journal 32, 20342046. Dixon, A., Nijemeisland, M., 2001. CFD as a design tool for xed-bed reactors. Industrial & Engineering Chemistry Research 40, 52465254. Germain, J., del Valle, J., de la Fuente, J., 2005. Natural convection retards supercritical CO2 extraction of essential oils and lipids from vegetable substrates. Industrial & Engineering Chemistry Research 44, 28792886. Guardo, A., Coussirat, M., Larrayoz, M.A., Recasens, F., Egusquiza, E., 2004. CFD ow and heat transfer in nonregular packings for xed bed equipment design. Industrial & Engineering Chemistry Research 43, 70497056.

A. Guardo et al. / Chemical Engineering Science 61 (2006) 4341 4353 Guardo, A., Coussirat, M., Larrayoz, M.A., Recasens, F., Egusquiza, E., 2005. Inuence of the turbulence model in CFD modeling of wall-to-uid heat transfer in packed beds. Chemical Engineering Science 60, 17331742. Gunjal, P., Ranade, V., Chaudhari, R., 2005. Computational study of a singlephase ow in packed beds of spheres. A.I.Ch.E. Journal 51, 365378. Harris, C.K., Roekaerts, D., Rosendal, F.J.J., Buitendijk, F.G.J., Daskopoulos, Ph., Vreenegoor, A.J.N., Wang, H., 1996. Computational uid dynamics for chemical reactor engineering. Chemical Engineering Science 51, 15691594. Hill, S., 1952. Channeling in packed columns. Chemical Engineering Science 1, 247253. Homsy, G.M., 1987. Viscous ngering in porous media. Annual Reviews of Fluid Mechanics 19, 271311. Jolls, K.R., Hanratty, T.J., 1966. Transition to turbulence for ow through a dumped bed of spheres. Chemical Engineering Science 21, 11851190. Joshi, J., Ranade, V., 2003. Computational uid dynamics for designing process equipment: expectations, current status, and path forward. Industrial & Engineering Chemistry Research 42, 11151128. Kays, W.M., 1994. Turbulent Prandtl number. Where are we? Journal of Heat Transfer 116, 284295. Kuipers, J.A.M., Van Swaaij, W.P.M., 1998. Computational uid dynamics applied to chemical reaction engineering. In: Wei, J., et al., Anderson, J., Bischoff, K., Denn, M., Seineld, J., Stephanopoulos, G. (Eds.), Advances in Chemical Engineering, vol. 24. Academic Press, San Diego, CA, pp. 227328. Lakshminarayana, B., 1991. An assessment of computational uid dynamic techniques in the analysis and design of turbomachinery. The 1990 Freeman scholar lecture. Journal of Fluids Engineering 113, 315352. Lloyd, B., Boehm, R., 1994. Flow and heat transfer around a linear array of spheres. Numerical Heat Transfer Part AApplications 26, 237252. Logtenberg, S.A., Nijemeisland, M., Dixon, A.G., 1999. Computational uid dynamics simulations of uid ow and heat transfer at the wall-particle contact points in a xed-bed reactor. Chemical Engineering Science 54, 24332439. Magnico, P., 2003. Hydrodynamics and transport properties of packed beds in small tube-to-sphere diameter ratio: pore scale simulation using an Eulerian and a Lagrangian approach. Chemical Engineering Science 58, 50055024. Manickam, O., Homsy, G.M., 1995. Fingering instabilities in vertical miscible displacement ows through porous media. Journal of Fluid Mechanics 288, 75102. Manz, B., Gladden, L.F., Warren, P.B., 1999. Flow and dispersion in porous media: LatticeBoltzmann and NMR studies. A.I.Ch.E. Journal 45, 18451854. Nijemeisland, M., 2000. M.Sc. Thesis. Worcester Polytechnic Institute. Worcester, MA, USA. Nijemeisland, M., Dixon, A.G., 2004. CFD study of uid ow and wall heat transfer in a xed bed of spheres. A.I.Ch.E. Journal 50, 906921. Oka, Y., Koshizuka, S., Jevremovic, T., Okano, Y., 1994. Systems design of direct-cycle supercritical-water-cooled fast reactors. Nuclear Technology 109, 110. Poliakoff, M., George, M.W., Howdle, S.M., 1996. Inorganic and related chemical reactions in supercritical uids. In: Van Eldik, R. (Ed.), Chemistry Under Extreme and Non-classical Conditions. Spektrum, Heidelberg, pp. 189218 (Chapter 5).

4353

Poling, B., Prausnitz, J., OConnell, J., 2000. The Properties of Gases and Liquids. McGraw-Hill, New York, pp. 3.110.56. Ramirez, E., Recasens, F., Fernandez, M., Larrayoz, M.A., 2004. Sunower oil hydrogenation on Pd/C in SC propane in a continuous recycle reactor. A.I.Ch.E. Journal 50, 15451555. Ranade, V., 2002. Computational Flow Modeling for Chemical Reactor Engineering. Academic press, New York, pp. 244422. Ranz, W.E., Marshall Jr., W.R., 1952. Evaporation from drops, part 1. Chemical Engineering Progress 48, 173180. Reid, R., Prausnitz, J., Poling, B., 1987. The properties of Gases and Liquids. McGraw-Hill, Boston, pp. 95205. Romkes, S.J.P., Dautzenberg, F.M., van den Bleek, C.M., Calis, H.P.A., 2003. CFD modeling and experimental validation of particle-to-uid mass and heat transfer in a packed bed at very low channel to particle diameter ratio. Chemical Engineering Journal 96, 313. Saiz, S., Larrayoz, M.A., Trabelsi, F., Recasens, F., 2000. Procedimiento para la extraccin de lanolina de la lana y planta para la realizacin de dicho procedimiento. Spain patent No. P200002461. Shent, J., Kaguei, S., Wakao, N., 1981. Measurements of particle-to-gas heat transfer coefcients from one-shot thermal responses in packed beds. Chemical Engineering Science 36, 12831286. Shur, M., Spalart, P., Squires, K., Strelets, M., Travin, A., 2005. Three dimensionality in Reynolds-averaged NavierStokes solutions around twodimensional geometries. AIAA Journal 43, 12301242. Sovov, H., Kuera, J., Je, J., 1994. Rate of the vegetable oil extraction with supercritical CO2 II. Extraction of grape oil. Chemical Engineering Science 49, 415420. Spalart, P.R., 2000. Strategies for turbulence modelling and simulations. International Journal of Heat and Fluid Flow 21, 252263. Stber, F., Vzquez, A.M., Larrayoz, M.A., Recasens, F., 1996. Supercritical uid extraction of packed beds: external mass transfer in upow and downow operation. Industrial & Engineering Chemistry Research 35, 36183628. Suekane, T., Yokouchi, Y., Hirai, S., 2003. Inertial ow structures in a simplepacked bed of spheres. A.I.Ch.E. Journal 49, 1017. Vortmeyer, D., Schuster, J., 1983. Evaluation of steady ow proles in rectangular and circular packed beds by a variational method. Chemical Engineering Science 38, 16911699. Wakao, N., 1976. Particle-to-uid transfer coefcients and uid diffusivities at low ow rate in packed beds. Chemical Engineering Science 31, 11151122. Wakao, N., Kaguei, S., 1982. Heat and Mass Transfer in Packed Beds. McGraw-Hill, New York, pp. 264295. Wakao, N., Kaguei, S., Funazkri, T., 1979. Effect of uid dispersion coefcients on particle-to-uid heat transfer coefcients in packed beds: correlation of Nusselt numbers. Chemical Engineering Science 34, 325336. White, F., 1991. Viscous Fluid Flow. McGraw-Hill, New York, pp. 482542. Yaws, C., 1999. Chemical Properties Handbook. McGraw-Hill, New York, pp. 155. Yin, F., Wang, Z., Afacan, A., Nandakumar, K., Chuang, K.T., 2000. Experimental studies of liquid ow maldistribution in a random packed column. Canadian Journal of Chemical Engineering 78, 449457.

S-ar putea să vă placă și