Você está na página 1de 13

Chemical Engineering Science 122 (2015) 197209

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Detailed numerical simulations of catalytic xed-bed reactors:


Heterogeneous dry reforming of methane
Gregor D. Wehinger n, Thomas Eppinger, Matthias Kraume
Chemical and Process Engineering, Technische Universitt Berlin, Fraunhoferstr. 33-36, 10587 Berlin, Germany

H I G H L I G H T S






Fully 3D CFD modeling of randomly packed catalytic xed-bed reactor.


CFD model combines complex geometry with detailed kinetics of DRM.
Meshing recommendations are given due to grid sensitivity studies.
Determination of regions with catalyst deactivation by surface-adsorbed carbon.

art ic l e i nf o

a b s t r a c t

Article history:
Received 23 May 2014
Received in revised form
2 September 2014
Accepted 4 September 2014
Available online 7 October 2014

Highly endothermic (or exothermic) heterogeneous catalytic reactions are performed commonly in
xed-bed reactors with small tube-to-particle-diameter ratios N both in industrial and lab-scale
applications. For these reactor congurations conventional plug ow models and pseudohomogeneous kinetic models fail. An adequate modeling can be carried out with full computational
uid dynamics (CFD) in combination with detailed reaction mechanisms. In this study, a full threedimensional xed-bed reactor for the catalytic dry reforming of methane (DRM) over rhodium was
simulated with a detailed reaction mechanism. The bed consists of 113 spherical solid particles in which
thermal conductivity was considered. Two different Reynolds numbers were investigated, i.e., 35 and
700. The simulated DRM xed-bed reactor demonstrates the strong interaction between chemical
kinetics and transport of momentum, heat and mass. The observed velocity, temperature and species
elds are characterized by their three-dimensional behavior and interactions highlighting their
complexity and discrepancy from lumped model predictions. In addition, the reaction mechanism
determines regions with catalyst deactivation by carbon deposition. This study demonstrates the
advantages of modeling heterogeneous catalytic xed-bed reactors with small N fully in threedimensional in combination with detailed reaction mechanisms. Finally, this modeling approach reduces
dependencies on empiricism for the calculation of multiscale reaction devices.
& 2014 Elsevier Ltd. All rights reserved.

Keywords:
CFD
Catalysis
Dry reforming of methane
Fixed-bed reactors
Modeling

1. Introduction
The atmospheric concentration of greenhouse gases, i.e.,
carbon dioxide (CO2), nitrous oxide (NO), methane (CH4) and
chlorouorocarbons (CFCs), has increased dramatically during
the last decades (Hartmann et al., 2013). These anthropogenic
emissions have risen a global concern over the current technological practices. Hence, the eld of interest involves CH4 and CO2
disposal, utilization and removal, as well as the effect of these
gases in the atmosphere (Mikkelsen et al., 2010; Centi and
Perathoner, 2009; Hunt et al., 2010; Papadopoulou et al., 2012;
n

Corresponding author. Tel.: 49 30 314 28733; fax: 49 30 314 21134.


E-mail address: gregor.wehinger@tu-berlin.de (G.D. Wehinger).

http://dx.doi.org/10.1016/j.ces.2014.09.007
0009-2509/& 2014 Elsevier Ltd. All rights reserved.

Balat and z, 2007). However, the general engineering interest lies


in processes in which CH4 and CO2 react to synthesis gas or syngas,
i.e., COH2. On one hand syngas can be used to produce liquid
fuels via the FischerTropsch reaction; a review given in Dry
(2002). On the other hand syngas can be utilized in chemical
energy transmission systems (Wang et al., 1996).
Dry reforming of methane (DRM) is such a process in which
CH4 and CO2 react to syngas:
CH4 CO2 2H2 2CO;

H  260 kJ=mol

This highly endothermic process is performed at temperatures of


7001000 1C. One of the largest obstacles for the industrial use of
DRM is coke formation, which quickly leads to a deactivation of the
catalyst (Chen et al., 2001; Ginsburg et al., 2005; Guo et al., 2007).

198

G.D. Wehinger et al. / Chemical Engineering Science 122 (2015) 197209

This prospective process has to be carried out with an appropriate


catalyst. In the last decades Nickel-based catalysts and noble metalsupported catalysts (Rh, Ru, Pd, Pt, Ir) have shown encouraging
performances regarding conversion and selectivity (Torniainen et al.,
1994; Wang et al., 1996; Guo et al., 2004). Rhodium for example is
characterized by its low afnity for carbon formation and its high
activity (Rostrup-Nielsen and Hansen, 1993; Bradford and Vannice,
1999).
Typical reactor types for endothermic (or exothermic) reactions
are xed beds, foams, multi-channel reactors or uidized-bed
reactors. Still today, the most common way to carry out a heterogeneous catalytic reaction is a xed-bed reactor. Randomly distributed catalytic particles are the simplest type of such a reactor,
whereas particle diameters range from 2 to 10 mm (Eigenberger,
2008). For the DRM the heat transfer inside the reactor is one of the
major engineering issues. Thus, small reactor tubes are desirable.
Additionally, high gas velocities and reasonable pressure drops
constrain the particle diameter to be quite large. Hence, xed beds
with a small tube-to-particle-diameter ratio 4 o N o 15 are preferable. In some lab-scale applications even N o4 were carried out, e.g.,
Leva et al. (1951), Reichelt (1972), Vortmeyer and Winter (1984), and
Dixon (1997). In all cases, randomly packed beds are characterized by
inhomogeneous structures. Especially for small N the inhomogeneities become dominant resulting in signicant wall effects, local back
ows and large axial as well as radial gradients. Consequently,
conventional descriptions, based on plug ow and pseudohomogeneous kinetics, are questionable for these xed-bed congurations (Bey and Eigenberger, 1997; Dixon, 1997; Bauer and Adler,
2003; Freund et al., 2005; Grf et al., 2014). The strong interplay
between velocity, temperature and species distribution makes the
xed-bed reactor a very interesting and likewise challenging device
for chemical engineers. It includes several time and length scales. The
multiscale modeling, or rst-principles approach (Dudukovic, 2009),
pursues to describe entirely the system by theory of the actual
phenomena, i.e., elementary reaction steps at the catalyst surface and
a detailed characterization of the uid ow. As a consequence, an
adequate description of catalytic xed-beds should include the
rigorous modeling with full computational uid dynamics (CFD) in
the interstitial regions as well as in the solid in combination with
detailed chemical reaction models (also called the micro-kinetics).
In recent years, several authors have simulated spatially
resolved xed-bed reactors accounting for radial, axial and circumferential proles (Dixon and Nijemeisland, 2001; Guardo et al.,
2005; Ookawara et al., 2007; Bai et al., 2009; Eppinger et al., 2011;
Behnam et al., 2013). However, only few have coupled the uid
dynamics of xed-beds with catalytic reactions (Kuroki et al.,
2009; Taskin et al., 2010; Dixon et al., 2012). That means the
already complex equation system will be extended by species
conservation equations. Several authors used pseudo-homogeneous kinetics in combination with detailed uid dynamics,
due to the small number of reaction equations. However, these
kinetics are often limited to a certain range of process parameters
and could therefore not be applied to other ow regimes or reactor
types (Salciccioli et al., 2011). Especially the species development
inside xed-bed reactors are often insufcient reproduced with
such kinetics in contrast to the exit concentrations, which was
recently demonstrated by Korup et al. (2011). As Wehinger et al.
(2014) concluded spatially resolved uid dynamics must be
combined with reliable kinetics, i.e., detailed reaction mechanisms. Several detailed methane reforming kinetics over rhodium
were published validated over an operating range relevant to
industrial applications and claimed to be thermodynamically
consistent (Hickman and Schmidt, 1993; Mhadeshwar and
Vlachos, 2005; Maestri et al., 2008, 2009; McGuire et al., 2011;
Kahle et al., 2013). Finally, the combination of a rst-principle
approach at different scales, i.e., detailed reaction mechanism at the

catalyst scale and full NavierStokes equations at the reactor scale,


helps to obtain a fundamental understanding of chemical reactors.
In this study, we investigated spatially resolved heterogeneous
catalysis of the dry reforming of methane over rhodium in a xedbed reactor in terms of combining full CFD simulations with a
detailed reaction mechanism from McGuire et al. (2011). A
catalytic xed-bed reactor with spherical solid particles and a
small tube-to-particle-diameter ratio (N 4) was numerically
simulated. The aim of the study was rstly to obtain a better
understanding of the strong interactions between catalytic reactions and the surrounding ow in xed-bed reactors. Secondly, we
investigated the feasibility to model catalytic xed-bed reactors in
an adequate multiscale way.

2. Simulating heterogeneous xed-bed reactors


2.1. Modeling chemically reactive ow
2.1.1. Governing equations
For the simulations in this study, full three-dimensional governing equations were applied. The conservation of total mass,
momentum in x-, y-, z-directions, mass of species and energy leads
to the solution for velocity, pressure, temperature and species
concentration in the calculation domain. A laminar problem with
Einstein convention can be written as follows.
Conservation of mass:
vi

0
xi
t

where is the mass density, t is the time, xi are the Cartesian


coordinates and vi are the velocity components.
Conservation of momentum:
vi vi vj p ij


g i
t
xi xj
xj

The stress tensor ij is written as follows:



  
v v
2
v
ij  i j ij k
3
xj xi
xk
where is the mixture viscosity and
which is unity for ij, else zero.
Conservation of species i:
Y i vj Y i ji;j

0
t
xj
xj

for

ij is the Kronecker delta,

i 1; ; Ng

with mass fraction Y i mi =m of species i and total mass m. Ng is


the number of gas phase species. The components ji;j of the
diffusion mass ux are described by the mixture-average formulation:
Y
X i DTi T
ji;j  i DM

i
Xi
xj T xj

where DM
i is the effective diffusivity between species i and the
remaining mixture. Xi represents the molar fraction of species i. Mi
is the molecular weight of species i and T is the temperature. The
binary diffusion coefcients Di are obtained through polynomial
ts. The molar fraction Xi can be written as
Xi

1
N

Yi
Y

j g 1 Mjj M i

Conservation of energy in terms of specic enthalpy h is as


follows:
vj
h vj h jq;j p
p

vj  jk
Sh
t
t
xj
xj
xk
xj

G.D. Wehinger et al. / Chemical Engineering Science 122 (2015) 197209

where Sh is the heat source. Diffusive heat transport jq;j is given by


the following:
Ng
T
jq;j  hi ji;j
xj i 1

with thermal conductivity of the mixture


enthalpy h
Ng

h Y i hi T
i1

and mixture specic


10

as a function of temperature hi hi T.
Ideal gas was assumed connecting pressure, temperature and
density to close the governing equations:
p

RT
N
i g 1 X i M i

11

199

interpreted as being uniform, cf. Deutschmann (2008) and Kee


et al. (2003). The molar net production s_ i can be written in the
following form:
Ks

Ng Ns

k1

j1

s_ i ik kf k

cj jk

17

where Ks is the number of surface reactions, cj are the species


concentrations, in mol/m2 for the adsorbed species Ns and in mol/
m3 for the gas phase species Ng, respectively. In addition, the
surface coverage takes into account the surface site density
(mol/m2), representing the maximum number of species that can
adsorb on a unit surface area. Furthermore, a coordinate number i
expresses the number of surface sites which are covered by this
species:

i c i i  1

18

i can be written as

Additionally, NASA polynomial functions were applied to derive


heat capacities cp;i .

the time-dependent variation of

2.1.2. Modeling turbulence


With the help of the particle Reynolds number Rep the ow in
xed-bed reactors can be characterized as follows:

Under steady-state conditions the left side of Eq. (19) will be zero.
The reaction rate expression can be modied by the concentration,
or coverage, of some surface species:




i
 i k i
 E ak N s
i k exp
20
kf k Ak T k exp
RT
RT
i1

Rep

vin dp

12

For Rep 4 300 the ow behaves highly unsteady, chaotic and


qualitatively resembling turbulent, cf. Zikowska and Zikowski
(1988). Consequently, such ow congurations were modeled
with the Reynolds-averaged turbulence approximation (RANS).
The solution variables are split into mean components v i and
uctuating components v0i . They are then integrated over an
interval of time much larger than the small-scale uctuations.
The turbulence momentum equation can then be written as
 
 p
v i


vi vj v0i v0j ij g i
13
xj
xi xj
t


The Reynolds stresses v0i v0j have to be put in terms of the
averaged ow quantities to close the system of equations. In our
case we used the realizable k model, developed by Shih et al.
(1995), in combination with a two-layer all-y wall treatment
driven by shear (Wolfshtein, 1969), cf. manual of STAR-CCM (CDadapco, 2014).
2.1.3. CFD and heterogeneous chemical reactions
The chemical reactions at the catalytic surface are coupled via
boundary conditions with the species distribution equation (5).
Under steady-state conditions gas-phase molecules of species
i, which are produced/consumed at the catalytic surface by
desorption/adsorption, have to diffuse from/to the catalyst
(Deutschmann, 2008):
 
! !
n ji Rhet
14
i
!
with the outward-pointing unit vector normal to the surface n
!
and the diffusion mass ux ji as in Eq. (6). The heterogeneous
reaction term Rhet
is given by
i
F cat=geo M i s_ i
Rhet
i

15

where Mi is the molar weight, s_ i is the molar net production rate of


gas-phase species i and F cat=geo is the ratio of catalytic active area
Acatalytic to geometric area Ageometric
F cat=geo Acatalytic =Ageometric

16

In all the simulations the mean-eld approximation was


applied to model the surface reactions. The assumption is that
adsorbates are randomly distributed on the surface, which is

s_ i i

19

with two extra coverage parameters, ik and ik. The term including ik indicates the modication of the surface rate expression
proportional to any arbitrary power of a surface species concentration. ik represents a modication of the activation energy as a
function of coverage.
The occurrence of adsorption reactions results in a modication
of the conventional rate coefcient by referencing sticking coefcients Si
s
S0
RT
ads
kf k i
21
2 M i
with S0i as the initial (uncovered surface) sticking coefcient and
Nj s 1 0jk is the sum of all the surface reactant's stoichiometric
coefcients, cf. Kee et al. (2003) and Deutschmann (2008).
Additionally, the operator splitting algorithm was implemented. The algorithm decouples the general species transport equation, due to the different time scales of the ow eld and the
chemical reactions. The time integration of the chemical state
(species mass fractions and enthalpy) was performed in two steps.
For the rst step, only the species source terms were taken into
account for the integration of a time interval. For the second, the
ow eld was integrated over a time interval without the chemical
source terms, cf. Ren and Pope (2008).
All simulations were realized with the simulation software
STAR-CCM version 9.02.005 by CD-adapco (2014). The equation
system for the surface species was solved by DARS, an add-in
solver for chemical reactions for STAR-CCM . The computational
time was high due to the mesh size and chemical reaction steps.
The laminar case with a 3.2 million cell mesh (M3) ran for 35,000
iterations which yielded in a total CPU time of 1:7  107 s or
196 days on a Intel Xeon 3.07 GHz CPU. The turbulent case with a
3.6 million cell mesh (M5) ran for 9000 iterations, i.e., 9:3  106 s
or 107 days. However, the simulations were performed on several
parallel CPUs, which reduced the computational time signicantly.
2.2. Detailed reaction mechanism
As already mentioned detailed uid dynamics call for detailed
reaction mechanisms. Still today, in many CFD simulations LangmuirHinshelwoodHougonWatson (LHHW) models are applied.

200

G.D. Wehinger et al. / Chemical Engineering Science 122 (2015) 197209

Table 1
Detailed surface mechanism for the dry reforming of methane, from McGuire et al.
(2011).
No.

Reaction

H2

- Hn Hn

O2

- On On

CH4

- CHn4

H2O

- H2On

CO2

- COn2

CO

Hn Hn - H2

On On - O2

- COn
n

A (a)

E (kJ/mol)

1:0  10  2 b

0.0

1:0  10  2 b

0.0

8:0  10  3 b

0.0

1:0  10  3 b

0.0

4:8  10  2 b

0.0

5:0  10  1 b

0.0

3:0  1021

77.8

1:3  1022

355.2
 280c

nO
9

H2On - H2O

10

COn - CO

3:0  10

13

3:5  1013

133.4

4:1  1011

18.0

14

25.1

5:0  1022

83.7

20

37.7

3:0  1020

33.5

5:0  1022

106.4

3:0  1021

100.8

3:0  1021

171.8

5:2  1023

97.9

2:5  1021

169.0

5:5  1018

121.6

3:0  1021

171.8

5:0  1019

108.9

3:7  1021

0.0

 15c

CO

11

COn2 - CO2

12

CHn4 - CH4

13

Hn On - OHn

14

OHn

15

Hn OHn - H2On

16

H2On

17

OHn OHn- H2On On

18

H2On On - OHn OHn

19

Cn O - COn

20

COn

21

COn On - COn2

22

COn2 n- COn On

23

COn Hn - HCOn

24

HCOn

nCO

45.0

1:9  10

- Hn On

- Hn OHn

- C n On
n

- COn Hn

3:0  10

50c
n

25

HCO

26

CHn On - HCOn

27

CHn4

28

CHn3 Hn - CHn4

29

CHn3

30

CHn2 Hn - CHn3

31

CHn2

32

CHn Hn - CHn2

33

CHn

34

Cn Hn - CHn

35

CHn4 On - CHn3 OHn

36

CHn3 OHn - CHn4 On

37

CHn3 On - CHn2 OHn

38

CHn2 OHn - CHn3 On

39

CHn2 On - CHn OHn

40

CHn OHn - CHn2 On

41

CHn On - Cn OHn

42

Cn OHn - CHn On

- CH O

n
n

- CHn3 Hn
n

- CHn2 Hn
n

- CHn Hn
n

- Cn Hn
n

3:7  10

24

59.5

3:7  1021

167.5

3:7  1021

61.0

3:7  1021

51.0

3:7  1024

103.0

3:7  1023

44.1

3:7  1024

100.0

3:7  1021

68.0

3:7  1021

21.0

3:7  1021

172.8

1:7  1024

80.34

21

24.27

3:7  10

3:7  1024

120.31

3:7  1021

15.06

3:7  1024

114.5

3:7  1021

36.82

3:7  1021

30.13

3:7  1021

136.0

Surface site density 2:72  10  9 mol/cm2.


a

Arrhenius parameters for rate constants. Units: pre-exponential factor A for


unimolecular reactions (s  1), for bimolecular reactions (cm2 mol  1 s  1).
b
Initial sticking coefcient S0i ().
c
Coverage dependent activation energy in Eq. (20). For more information see
e.g., Kee et al. (2003).

However, the fundamental mechanism described by LHHW models can be elusive and the physical signicance of parameters can
be seriously questionable, cf. Salciccioli et al. (2011) and Wehinger
et al. (2014). Taking the strong interactions between the different
transport quantities into account an erroneous kinetics will
directly lead to misleading predictions. Therefore, it is recommended only to use reliable kinetics for detailed uid dynamics
simulations.
In this study a detailed reaction mechanism, see Table 1, was
implemented published by McGuire et al. (2011) for the DRM over

rhodium supported strontium-substituted hexaaluminate. This


kinetic model describes elementary-like processes occurring at
the catalytic surface with the mean-eld approximation distinguishing between adsorption, surface reaction and desorption. The
mechanism was originally developed for the catalytic partial
oxidation of iso-octane on rhodium published by Hartmann et al.
(2010). Several kinetic parameters were then modied to describe
the DRM adequately. The surface reaction mechanism consists
of 42 irreversible elementary reactions including 12 surfaceadsorbed species and 6 gas-phase species, as well as reactions
involving HCOn. In Table 1 the asterisk represents a surface site or
surface adsorbed species. The detailed DRM mechanism was
previously validated by a three-dimensional calculation of the
experiments which were carried out in a stagnation ow reactor,
cf. Wehinger et al. (2014).
2.3. Generation of random packings and meshing
The scheme of the spherical xed-bed reactor is shown in Fig. 1. In
addition to the xed bed an upstream and downstream region was
generated to minimize the inuence of the boundary conditions. The
geometric quantities of the reactor are the following: catalytic bed
height H40 mm, reactor diameter D16.2 mm, sphere diam
eter dP 4.09 mm, which leads to a tube-to-particle-diameter ratio
N D=dP 3:96. The reactor contains 113 spheres, which were packed
randomly. This leads to a specic particle area a AP;total =V Reactor of
784 m2/m3. The geometric generation of the randomly packed bed was
carried out with a discrete element method (DEM) and is described in
detail elsewhere (Eppinger et al., 2011; Zobel et al., 2012). In the DEM
simulation the tube was lled up with particles. When all particles were
settled, the geometric information of the particle centroids was
extracted. This information was then used to build up the desired
packed-bed.
A polyhedral grid was chosen to mesh the solid particles and the
gas phase. Additionally, prism layers were introduced at the interface
between uid and solid phases. In the meshing process the particles
were attened locally at particlewall contact-points and particle
particle contact-points to avoid bad cell qualities, see Fig. 2. Besides
attening the spheres at contact points, several other methods exist,

Inlet

Upstream
h u = 1/4 H

Flow direction
Spheres
d p = 4.09 mm

Fixed-bed
H = 40 mm

Reactor diameter
D = 16.2 mm
Downstream
h d = 1/2 H

Outlet
Fig. 1. Scheme of the spherical xed-bed reactor.

G.D. Wehinger et al. / Chemical Engineering Science 122 (2015) 197209

201

diameter dp 4:09 mm. On the catalytic surface the detailed


mechanism of the DRM was implemented, see Table 1. Likewise
as in the experiments by McGuire et al. (2011) processes in the
pores were treated as instantaneous diffusion resulting in an
enlarged catalytic active area F cat=geo 90 in Eq. (15). Furthermore,
a constant reactor-wall temperature and inlet temperature
T wall T inlet 973 K was chosen. The reactor was in steady-state
operation at ambient pressure, which is indicated by the pressure
outlet pout 1 bar. The spheres were treated as solid particles with
conjugate heat transfer that means the temperature of the solid
was not constant. The solid density cat was set to 2214 kg/m3,
specic heat cp;cat 850 J/(kg K) and thermal conductivity
cat 12:6 W/(m K), as reported in Schwiedernoch et al. (2003)
for alumina monolith including porosity.

3. Results and discussion


3.1. Porosity

Fig. 2. Section of xed-bed reactor meshing (A) mesh M3 for Rep 35 and (B) mesh
M5 for Rep 700. Gas phase polyhedral mesh in gray and solid sphere polyhedral
mesh in dark. Flattening is visible between the particles.

Table 2
Characteristics for the investigated meshes.
Mesh Prism layer
thickness (mm)

No. of
layers

Base size
(mm)

Total no. of
cells [106]

Laminar Re 35
M1
0.19
M2
0.19
M3
0.19

2
2
2

8
4
3

0.6
2.6
3.2

0.11
0.65
0.67

Turbulent Re 700
M4
0.0427
M5
0.0427
M6
0.0427

2
2
2

4
3
2

2.8
3.5
10.5

0.66
0.67
3.38

No. solid
cells [106]

as discussed in Dixon et al. (2013). In all the simulations the


boundary layer thickness BL was resolved with two or three prism
layers. BL was approximated with a correlation for the stagnation
point for spheres (Dhole et al., 2006):

BL
dp

1:13  Rep 0:5

22

In addition, as recommended by Dixon et al. (2013), the dimensionless cellwall distance y was kept to approx. 0.52.0. For the
investigated cases the velocity boundary layer (BL) thickness was
smaller than the temperature BL and the concentration BL. The
Prandtl number Pr was of the same order of magnitude as the
Schmidt number Sc. The inuence of the mesh size was investigated
by means of mesh renement, see Table 2. The base size is a
characteristic dimension of the mesh model, to which all other mesh
dimensions refer. It can be interpreted as a scaling factor of the mesh
resulting in the total number of cells (CD-adapco, 2014).
2.4. Boundary conditions
The conditions at the inlet were the following: feed gas
composition xin;CO2 =xin;CH4 =xin;N2 0:20=0:10=0:70, inlet velocity
vin;1 0:886 m=s or vin;2 17:72 m=s. The corresponding particle
Reynolds numbers were Rep 35; 700 calculated with a mean
dynamic viscosity gas;973 K 9:504  10  5 m2 =s and the particle

The local porosity as a function of dimensionless wall


distance was obtained by averaging the local porosity distribution in terms of height and circumference over 40 cylindrical
planes of different radii inside the xed bed. The rst and the last
layer of particles were not taken into account to avoid edge effects.
In Fig. 3 the porosity of the computer-generated bed is compared
with experimental data from Mueller (1992) and a more general
equation from de Klerk (2003):

r 2:14r 2 2:53r 1 if z r 0:637


r 0:29 exp 0:6r  cos 2:3 r  0:16
0:15 exp  0:9r

if z 4 0:637

23

The rst peak of the porosity  1 occurs at the wall 0, where


spheres touch the wall in contact points. The second peak at 1
can be reproduced by the simulation. In the center of the bed, 2,
experimental values are higher than for the virtual bed. The minimum values of the experiments are lower than the computergenerated packing and they occur at smaller wall distances. The
more general equation from de Klerk (2003) is valid for a wide range
of N and therefore not too accurate to compare with the simulated
case with N  4. However, the trend is comparable. The reason for
the discrepancy between simulation and experiments is the loose
packing structure, as it can be seen in Fig. 1, and the low height-toradius ratio of the computer-generated bed H=D  2:47 in comparison with the experiments from Mueller (1992), H=D  7:84. Consequently, inhomogeneities are more dominant toward the averaged
results. For loose packings the porosity is generally higher than for
more dense packings, except for the center of the bed. That means in
1
Computer-generated packing
Mueller (1992)
de Klerk (2003)

0.8
0.6
0.4
0.2
0

0.5

1.5

= (R-r)/dp (-)
Fig. 3. Comparison of porosity as a function of dimensionless wall distance
R  r=dp between computer-generated packing and experimental measurements from Mueller (1992) averaged over the reactor height and a general equation
from de Klerk (2003).

202

G.D. Wehinger et al. / Chemical Engineering Science 122 (2015) 197209

the simulation the channeling in the center is not as severe as in the


experiments. Such packing effects have a great inuence on the local
porosity and, therefore, on the uid behavior, e.g., pressure drop and
velocity distribution. The overall porosity in the computer-generated
bed accounts to 0.5, whereas in Mueller's experiments a global
porosity of 0.47 was measured. That highlights the loose packing
structure of the simulated bed. Surface roughness and lling speed
inuence the packing density. Mueller (1992) for example used
plexiglas polished spheres to obtain radial porosity distributions. In
contrary, porous spheres, which have a rougher surface, would have
a looser packing density. Nonetheless, Eppinger et al. (2011) and
Zobel et al. (2012) demonstrated good accuracy between simulated
and experimental packed-bed quantities, i.e., local porosity and
pressure drop, with the same DEM-bed generation procedure. That
gives us condence that this method can be used for simulating
heterogeneous catalytic packed-beds. However, lling speed and
surface roughness should be aligned with an industrial relevant case.

3.2. Pressure drop


The pressure drop in a xed-bed reactor is often a crucial
parameter, since it designates the necessary energy for pumps and
compressors. The Ergun equation (Ergun, 1952) describes the
pressure drop adequately for innite packed beds:

!
150 1  2
1 
H
1:75
  v2in

3
Rep 3
d
p

24

However, it shows discrepancies when conning walls inuence the


ow eld, i.e., for small D=dp ratios. Several authors highlighted that
the effect is Reynolds number dependent. Therefore, Eisfeld and
Schnitzlein (2001) adapted Reichelt's equation (Reichelt, 1972)
paying special attention to tube-to-particle-diameter ratios D=dp ,

Pressure drop p (Pa)

10000

Eisfeld Eq.
Ergun Eq.
Simulation w/o reactions
Simulation w reactions

1000

T = 973 K
T = 873 K

100

10
10
100
1000
Particle Reynolds number ReP = vindP/

for spherical particles reading

154  A2w 1  2 Aw 1 

Rep
Bw 3
3

!


H
  v2in
dp

25

with the mean void fraction and coefcients Aw and Bw:


Aw 1
"

2
3  D=dp  1 


dp
Bw 1:15 
D

26
#2

2
0:87

27

In Fig. 4 and Table 3, the pressure drop as a function of particle


Reynolds numbers is given for the CFD simulations with and
without catalytic reactions for T 973 K and calculated with Eqs.
(24) and (25). The temperature decreases due to endothermic
reactions when DRM takes place, hence gas properties change.
Therefore, the denition of the Reynolds number is not explicit.
The Eisfeld equation as well as the Ergun equation was calculated
for two different reference temperatures, i.e., 873 K and 973 K,
with a feed gas mixture composition, resulting in different
viscosities and densities. For low N, the Ergun equation underestimates the pressure drop for low Rep and overestimates the
pressure drop for turbulent regimes, which was demonstrated as
early as Reichelt (1972). As it can be seen, the simulated pressure
drops are between the Ergun and Eisfeld equation for 873 K, i.e.,
lower pressure for equivalent Rep . It has to be kept in mind that
the simulated bed has a low H/D ratio. Eisfeld's equation was
developed for much larger xed-beds. The low pressure drop in
the simulated bed might result from the loose packing and the low
H/D ratio leading to wall channeling and strong effects of the edge
zones of the bed, respectively. Several groups have validated their
simulations with pressure drop predictions and found good
agreement especially for low Re, e.g., Guardo et al. (2005) and
Freund et al. (2005). Concerning pressure drop grid independence
is reached with mesh renement M2 in the laminar case, see
Table 3. For the turbulent case, the pressure drop increases with
mesh renement. Though the deviation between smallest and
largest mesh is less than 5%. As it can be concluded, the simulated
pressure drops are in reasonable accuracy with predictions from
the literature even in the turbulent regime. However, local
quantities provide more information about mesh dependence.
Therefore, in Figs. 9 and 10 velocity, temperature and species
proles are compared at three different positions for the investigated meshes and discussed in the following chapters. The three
lines, i.e., in the stagnation zone above one of the rst spheres, in a
channel between a sphere and the wall, and in the interstitial area
between two spheres, are highlighted in Fig. 8.
3.3. Velocity distribution

Fig. 4. Pressure drop over particle Reynolds number. Comparison between simulation (meshes M3 and M5) and Eqs. (24) and (25), respectively.

!
Fig. 5 shows the specic velocity distribution j v j=vin on a plane
cut through the xed bed for Rep 35 M3 and Rep 700 M5.

Table 3
Results of investigated meshes for laminar and turbulent cases.
Mesh No.
Laminar Re 35
M1
M2
M3

pw=o;chem: (Pa)

34.7
37.9
37.9

Turbulent Re 700
M4
4218
M5
4220
M6
4400

pw;chem: (Pa)

33.2
36.5
36.5
4120
4124
4300

y ()

X CH4 (%)

X CO2 (%)

YCO (%)

YH2 (%)

Y H2O (%)

COn ()

C n ()

32.7
35.2
35.4

19.8
21.9
22.1

28.1
29.7
29.9

27.5
28.7
28.8

1.5
2.2
2.2

0.512
0.534
0.535

0.098
0.058
0.056

0.59
0.59
0.65

11.1
10.9
11.6

8.4
8.2
8.9

5.0
5.0
5.0

5.7
5.7
5.7

0.3
0.3
0.3

0.346
0.346
0.343

0.009
0.009
0.008

G.D. Wehinger et al. / Chemical Engineering Science 122 (2015) 197209

203

Fig. 6. Backow regions, i.e., cells with negative velocities. (A) for Rep 35 and
(B) for Rep 700.

2.5
!
Fig. 5. Specic velocity distribution j v j=vin on a plane cut through the xed bed.
(A) for Rep 35 M3 and (B) for Rep 700 M5.

2
(vz ())/vin (-)

The ow direction is from top to bottom. In both cases the ow


eld around the particles is highly three-dimensional. Axial as well
as radial differences occurs. Several different kinds of characteristic zones can be noticed: stagnation zones in front of particles,
wake and eddying behind particles, acceleration in void regions,
deceleration and channeling, especially in the near wall region.
!
The highest specic velocities j v j=vin  7 are found for the ow
eld of Rep 35, see Fig. 5(A). Thus, in the turbulent ow regime,
Fig. 5(B), the non-axial velocity components must be larger than
for the quasi-laminar ow. In Fig. 6 besides the spheres cells with
zero or negative velocities are illustrated. For Rep 700 these
regions are larger than for the laminar case. The ow is highly
characterized by back ow regions and non-axial velocity
components.
The local articial axial specic velocity vz =vin as a function
of dimensionless wall distance for different particle Reynolds
numbers is presented in Fig. 7. It follows the local porosity in
Fig. 3. High velocities are found in regions with high void fraction.
Close to the reactor wall the velocity decreases due to the
boundary layer and no-slip condition at the wall. The highest
articial velocities are in the range of 2.02.5, which was also
observed in experiments (Giese et al., 1998). The largest differences in the articial axial velocities for different particle Reynolds
numbers are found in the region close to the wall and in regions
with high void fractions. The different thicknesses of boundary
layers for laminar and turbulent regimes can be clearly seen.
Additionally, the diagram highlights that in the turbulent regime
radial and circumferential velocity components contribute to a
more leveled velocity distribution. It should be noted that back-

Rep = 35
Rep = 700

1.5
1
0.5
0

0.5

1.5

= ( R-r)/dp (-)
Fig. 7. Articial axial specic velocity vz =vin as a function of dimensionless wall
distance for different Reynolds numbers.

ow regions are not detected, due to averaging of the axial


velocities. Consequently, taking into consideration only twodimensional velocity distributions can be misleading.
In Figs. 9 and 10(a), (e) and (i), the specic velocity proles are
shown for the different meshes at three positions. The boundary
layer in the stagnation zone (a) is well resolved by all meshes.
However, the ow eld differs from single sphere proles due to
disturbance by other spheres, cf. Fig. 8. The channeling between
the wall and a sphere is shown in Figs. 9 and 10(e), i.e., position 2.
For the laminar case a parabolic velocity prole occurs with a
specic velocity of approx. 6. The peak is moved to the sphere's
side. For the turbulent case a typical prole is shown with a steep
gradient near the surfaces and a attened center with approx.
v=vin 4:5. Position 3 (i) represents the area between two spheres,
which is highly inuenced by the surrounding ow eld. Two
velocity peaks can be recognized located near the surfaces. The
ow decelerates in the center due to a recirculation zone further
upstream. In the laminar case the velocity increases smoothly,

204

G.D. Wehinger et al. / Chemical Engineering Science 122 (2015) 197209

Position
1

Position
2

Position
3

Fig. 8. Velocity magnitude contours for Rep 35 and positions for the mesh
validation.

whereas it shows a steep rise for the turbulent case. In Fig. 9(a) the
calculated velocities from mesh 1 are slightly different than for
mesh M2 and M3. On the contrary, the meshes M2 and M3 show
almost identical results in the laminar case. For higher Reynolds
numbers, only at position 3 the meshes show different velocities
with mesh renement. Here, the minimum velocity decreases in
the center with a ner mesh.
Finally, Fig. 11 shows the frequency of y values for meshes
M4M6. As it can be seen, most of the cells are small enough that
y o1:5. Hence, velocity boundary layers are well resolved.
3.4. Temperature distribution
The temperature distribution in the xed bed, i.e., gas phase and
solid particles, is shown in Fig. 12. The inlet temperature and the wall
temperature are set constant to 973 K. Due to the endothermic
reactions the temperature inside the bed decreases. Again, strong
axial and radial temperature differences up to approx. 80 K occur.
Low Reynolds numbers result in large residence times. Hence, in
Fig. 12(A) the overall temperature is lower than in Fig. 12(B). In (A), a
cold spot appears after approx. half of the reactor length, whereas in
(B) the temperature in the center decreases constantly. This is due to
the shorter residence time which moves the cold spot out of the bed.
In Fig. 12(B) the hot ow reaches deep inside the xed bed, whereas
for Rep 35 the ow cools down immediately. The solid particles can
be detected easily due to their almost constant temperature, which is
caused by the high thermal conductivity. As a result of channeling

in the near wall region, the thermal penetration into the bed is
declined. Again, the transport property temperature shows highly
three-dimensional behavior.
In Figs. 9 and 10(b), (f) and (j), temperature proles are shown
for the different meshes at the three positions. At position 1 the
temperature decreases from the inlet to the sphere's surface due to
endothermic reactions. However, for Rep 35 it is lower than for
the turbulent case. At position 2, i.e., between wall and sphere, the
temperature decreases from the constant wall temperature T 973
K to the specic surface temperature, which is inuenced rstly by
the surrounding ow and secondly by the surface reactions. In the
laminar case the temperature decreases almost linearly from the
wall to the surface. For the higher Reynolds number boundary
layers can be noticed near both surfaces, which are of the order of
magnitude of the velocity boundary layers. Finally, between the
spheres at position 3 (j) the temperature eld is highly inuenced
by the ow eld. In the laminar case the temperature decreases
from the outside to the center. Therefore, the left surface in (j) is
cooler than the right side. Again, an almost linear prole is shown
although the recirculation zone brings cooler gas. For Rep 700
the endothermic reactions cool down the surface, whereas hot gas
passes this position. The recirculation zone is larger than for lower
Reynolds. Hence, the temperature decreases strongly in the center.
3.5. Surface adsorbed species
As mentioned before, catalyst deactivation through carbon
deposition is one of the major draw backs of the DRM. It has to
be noticed that in reality coke formation takes place including
many carbonaceous atomic layers. However, the present reaction
mechanism only accounts for monolayer carbon on the surface.
Consequently, the model determines the regions where coking
takes place rather than the amount of coke. In Fig. 13 surface site
fractions of adsorbed carbon and some streamlines are illustrated
for different Reynolds numbers. As it can be seen, the carbon
deposition is not only dependent on the Reynolds number but is
due to the interactions between velocity, temperature and gas
composition. In Fig. 13(A) several regions of spheres are totally
blocked by carbon mainly in the center of the inlet region of the
bed. Hence, the catalyst is deactivated resulting in declined or
stopped production of syngas. In Fig. 13(B) almost no carbon is
adsorbed. Catalyst deactivation by carbon deposition for DRM
especially in the inlet regions of xed beds was recently observed
experimentally and numerically by Kahle et al. (2013). Fig. 13
highlights the advantage of this type of reaction mechanism for
DRM that can contribute to identify conditions and regions where
deactivation of the catalyst is likely to occur. In addition, Fig. 14
shows radially and circumferentially averaged surface site fractions of the adsorbed species Cn, COn, Hn and RHn. For the laminar
case (A), surface adsorbed carbon monoxide (COn) becomes the
most abundant reaction intermediate (MARI) after approx. 10 mm
in the xed-bed. Adsorbed carbon is only dominant in the
entrance of the reactor, whereas Hn occurs on less than 1% of the
surface sites. For the turbulent case (B), COn is again the MARI. Due
to the lower residence time, its surface fraction is lower, too. Cn
and Hn are found on less 2% on the surface. These two gures
illustrate that the DRM is kinetically limited. However, it has to be
kept in mind that the two cases are not under iso-conversion.
Therefore, a true comparison of location and quantity of surface
adsorbed species cannot be undertaken.

3.6. Gas phase species distribution


Radially and circumferentially averaged mole fractions of reactants and products as well as temperature are shown in Fig. 15

G.D. Wehinger et al. / Chemical Engineering Science 122 (2015) 197209

970

0.4
0.2

950
940

920
1

Distance from surface [mm]

980

970
Temperature [K]

7
6
5
4
3

0.5

1.5

0
0

Distance from surface [mm]

CO2

940
930

Wall

0.1
CH4
0.05

0.5

1.5

0.5

1.5

0.2

905

0.15

Radial distance between surfaces


[mm]

1.5

Radial distance between surfaces


[mm]

1.5

0.04

0.1

CH4

0.03
0.02
0.01

0
0.5

0.5

0.05

CO2

0.05

Radial distance surface to


wall [mm]

Mole fraction CO [-]

Mole fraction [-]

Temperature [K]

900

890
1.5

0.02

Radial distance surface to


wall [mm]

910

895

0.03

0.01

0.5

0.04

0.05

0.15

950

900

Wall

Radial distance surface to


wall [mm]

Distance from surface [mm]

0.2

960

Radial distance surface to


wall [mm]

0.02
0.01

Wall

0.03

910

1
0

0.1

920

2
Wall

Specific velocity v/vin [-]

Distance from surface [mm]

CH4

0
0

Mole fraction [-]

0.15

0.05

930

Specific velocity v/vin [-]

960

0.04
Mole fraction CO [-]

0.6

0.2

Mole fraction CO [-]

0.8

0.05
CO2

Mole fraction [-]

0.25

980

Temperature [K]

Specific velocity v/vin [-]

1.2

205

0
0

0.5

1.5

Radial distance between surfaces


[mm]

0.5

1.5

Radial distance between surfaces


[mm]

Fig. 9. Results of mesh renement for laminar case Rep 35. Specic velocity for (A) position 1, (E) position 2, and (I) position 3. Temperature for (B) position 1, (F) position 2,
(J) position 3. Mole fractions CO2 and CH4 for (C) position 1, (G) position 2, and (K) position 3. Mole fractions CO for (D) position 1, (H) position 2, and (L) position 3.

over the reactor length for different particle Reynolds numbers.


Catalytic conversion can be noticed and beside the main products
H2 and CO also water is formed. Water is the result of the reverse
watergas shift (WGS) reaction, CO2 H2 CO H2 O. Under
common DRM reforming conditions WGS is extremely rapid
(Rostrup-Nielsen and Hansen, 1993). For larger residence times
more hydrogen and carbon monoxide are produced. It becomes
clear that both reactors are economically not feasible, because only
few syngas is formed.
However, the DRM kinetics is highly inuenced by the reactor
temperature and likewise inuences it. This is strongly demonstrated in Fig. 16, where the mole fraction of H2 and surface site
fraction of carbon on a plane cut are shown. The strong interplay
between velocity and temperature distribution and the resulting
reactions can be seen. The low temperature and blockage of the
catalyst leads to a weak hydrogen production in the bed center in

Fig. 16(A). Whereas in stagnation zones, e.g., between spheres, the


production is high due to high residence time and low convection.
Rows three and four in Figs. 9 and 10 show mole fractions of
CO2, CH4 and CO in the gas phase and at surfaces. For the laminar
case, at position 1 no syngas is produced due to complete catalyst
blockage by Cn. On the contrary, for higher Reynolds numbers, CO
is produced and a boundary layer larger than the temperature BL
can be recognized, cf. Fig 10(d). At positions 2 and 3 the mole
fractions of methane and carbon dioxide decrease at the catalytic
surfaces, while syngas is produced. In the laminar case, mesh M1
shows lower conversion than M2 and M3. This could be due to the
lower discretization of the surface. Consequently, the velocity eld
as well as the temperature eld is affected. The meshes for the
turbulent case show in general similar species proles.
Comparing Figs. 9, 10, 15 and 16 it becomes clear that averaged
proles can be illusive, due to the fact that they neglect the radial

G.D. Wehinger et al. / Chemical Engineering Science 122 (2015) 197209

980

0.25

0.012
CO2

0.6
0.4

970
965
960
955
950

0.2

0.01

0.2
Mole fraction [-]

Temperature [K]

Specific velocity v/vin [-]

975
0.8

0.15
CH4
0.1

Mole fraction CO [-]

206

0.05

940
1

0
0

Distance from surface [mm]

Distance from surface [mm]

0.2 0.4 0.6 0.8

3
2

960
950

0.5

1.5

0.1
CH4

930

0.006
0.004

Wall

0.5

1.5

Radial distance surface to


wall [mm]

Radial distance surface to


wall [mm]

0.5

1.5

Radial distance surface to


wall [mm]

950

0.2

945

0.15

0.5

1.5

Radial distance surface to


wall [mm]

0.02

3
2

Mole fraction [-]

Temperature [K]

940

935

Mole fraction CO [-]

CO2

5
Specific velocity v/vin [-]

0.008

Wall

Wall

Wall

0.002

1
0

0.01

CO2

0.05

940

0.2 0.4 0.6 0.8

0.012

0.15
Mole fraction [-]

Temperature [K]

Specific velocity v/vin [-]

970

Distance from surface [mm]

0.2

6
5

Distance from surface [mm]

980

0.004

0
0

Mole fraction CO [-]

0.006

0.002

945
0

0.008

0.1
CH4
0.05

0.015

0.01

0.005

1
0

0.5

1.5

Radial distance between


surfaces [mm]

930

0.5

1.5

Radial distance between


surfaces [mm]

0.5

1.5

Radial distance between


surfaces [mm]

0.5

1.5

Radial distance between


surfaces [mm]

Fig. 10. Results of mesh renement for turbulent case Rep 700. Specic velocity for (A) position 1, (E) position 2, and (I) position 3. Temperature for (B) position 1,
(F) position 2, and (J) position 3. Mole fractions CO2 and CH4 for (C) position 1, (G) position 2, and (K) position 3. Mole fractions CO for (D) position 1, (H) position 2, and
(L) position 3.

and circumferential differences, e.g., boundary layers, of species


concentrations.

4. Conclusion
Highly endothermic (or exothermic) heterogeneous catalytic
reactions are performed mostly in xed-bed reactors with small
tube-to-particle-diameter ratios N. Inhomogeneities in the bed
structure are dominant especially for small N. This results in
signicant wall effects, local back ows and large axial and radial

gradients. For these reactor congurations conventional plug ow


models and pseudo-homogeneous kinetic models fail. An adequate modeling can be carried out with full CFD and detailed
reaction mechanisms.
In this study, a fully three-dimensional xed-bed lab-scale reactor
for the catalytic dry reforming of methane was simulated. A DEMmethod was applied to generate a randomly packed bed. The
meshing method takes into account boundary layers and particle
particle contact-points. The detailed DRM reaction mechanism
distinguishes between adsorption, surface reaction and desorption.
The bed consists of 113 spherical solid particles with applied

G.D. Wehinger et al. / Chemical Engineering Science 122 (2015) 197209

conjugate heat transfer. Two different Reynolds numbers were


investigated, i.e., Rep 35 and 700. Although the bed dimensions
are not large scale important ndings can be derived. The DRM xed
bed reactor demonstrates the strong interactions between chemical
kinetics and transport of momentum, heat and mass. The observed
velocity, temperature and species elds are characterized by their
0.06
M4, 2.8M cells
M5, 3.5M cells
M6, 10M cellss

three-dimensional behavior and interactions highlighting their complexity and discrepancy from lumped model predictions. Additionally, the reaction mechanism can detect regions where coking takes
place with the help of surface adsorbed carbon. We recommend
meshes with most of the near wall cells being small enough that
y o 1:5. This could be achieved by using two prism layers with a
total thickness calculated by Eq. (22). Meshes with approx. 3 million
total cells show grid independent results for laminar ows. However,
turbulent ows need ner meshes.
1

0.03

0.01
CO*

0.6

0.02
0.01
0

Rep = 35

H*

0.8

0.5

1.5

2.5

3.5

4.5

y+ [-]
Fig. 11. Frequency distribution over dimensionless wall distance y for different
meshes.

0.4

0.005
Rh*

0.2

C*

1
Rep = 700

0.8

0.01

H*

0.6

Rh*

0.4

Surface site fraction H* [-]

0.04

Surface site fraction [-]

Frequency [-]

0.05

207

0.005

CO*

0.2
C*

0.01

0.02

0
0.04

0.03

Reactor length z [m]


Fig. 14. Mean surface site fractions of Rhn, COn and Cn over reactor length. (A) for
Rep 35 mesh M3 and (B) for Rep 700 mesh M5.
Reacting zone

0.25

Rep = 35

960

CO2

0.15

Mole fraction [-]

0.1

940

Temp.
CH4

0.25

900

H2

H2O

880
Rep = 700

980
960

CO2

0.15
0.1

920

CO

0.05

0.2

980

Temp.
CH4

Temperature [K]

0.2

940
920

0.05

CO

H2O

0
-0.01 0 0.01

0.03

0.05

900
H2

880

Reactor length z [m]


Fig. 12. Temperature distribution on a plane cut through the xed bed. (A) for
Rep 35 mesh M3 and (B) for Rep 700 mesh M5.

Fig. 15. Mean mole fractions and mean temperature over reactor length. (A) for
Rep 35 mesh M3 and (B) for Rep 700 mesh M5.

Fig. 13. Catalyst deactivation through carbon deposition on the surface. (A) for Rep 35 mesh M3 and (B) for Rep 700 mesh M5.

208

G.D. Wehinger et al. / Chemical Engineering Science 122 (2015) 197209

Fig. 16. Hydrogen production and surface adsorbed carbon on a plane cut through the xed bed. (A) for Rep 35 mesh M3 and (B) for Rep 700 mesh M5.

The computational time of such detailed simulations is high,


i.e., several days on a cluster. Consequently, they are too expensive
and not practical for routine design of a xed-bed reactor. However, they are capable to support fundamental understanding of
the transport phenomena and kinetics. Therefore, such simulations can contribute to determine optimal reactor conditions. For
the next step, the model should be validated with spatially
resolved experimental data in a way recently carried out by the
Horn group, cf. Horn et al. (2010), Korup et al. (2011), and Geske
et al. (2013), or by the Paul Scherrer Institute in Switzerland (Bosco
and Vogel, 2006).
This study demonstrates the advantages of modeling heterogeneous catalytic xed-bed reactors with small N fully threedimensional in combination with reliable detailed reaction
mechanisms. In that way, resolved simulations can contribute to
a better understanding and therefore better choice of multiscale
chemical reactors. Finally, this modeling approach reduces dependencies on empiricism for the calculation of multiscale reaction
devices.

Pr
r
R
Ri
Rep
s_
Sc
Si
Sh
t
T
vin
vi
v0i
xi
Xi
y
Yi

Prandtl number Pr cp = ()
radial coordinate (m)
ideal gas constant (J/K mol)
production rate of species i (kg/m3 s)
particle Reynolds number Rep vin dp = ()
molar net production rate (mol/m3 s)
Schmidt number Sc =D ()
sticking coefcient ()
heat source (W/m3)
time (s)
temperature (K)
supercial velocity (m/s)
mean velocity components
uctuating velocity components
coordinate in i direction (m)
molar fraction of species i()
dimensionless distance from wall ()
mass fraction of species i ()

Greek letters
Nomenclature
Latin letters
a
Ak
c
cp
dp
D
Di
Ea
F cat=geo
h
H
ji
jq
k
k
m
Mi
N
Ng
p

specic particle area (m2/m3)


pre-exponential factor (s  1 or m2/mol s)
species concentration (mol/m3 or mol/m2)
specic heat capacity (J/kg K)
particle diameter (m)
tube diameter (m)
binary diffusion coefcient (m2/s)
activation energy (kJ/mol)
ratio of catalytic active area to geometric area ()
specic enthalpy (J/kg)
bed height (m)
diffusion mass ux (kg/m2 s)
diffusion heat transport (W/m2)
turbulent kinetic energy (J/kg s)
reaction rate constant (s  1 or m2/mol s)
mass (kg)
molar weight of species i (kg/mol)
tube-to-particle-diameter ratio ()
number of gas phase species ()
pressure (Pa)

BL
ij

surface site density (mol/m2)


boundary layer thickness (m)
Kronecker delta ()
parameter for modied activation energy
porosity ()
turbulent dissipation rate (J/kg s)
surface coverage ()
thermal conductivity (W/m K)
dynamic viscosity (Pa s)
parameter for modied surface rate expression
kinematic viscosity (m2/s)
dimensionless wall distance R  r=dp ()
uid density (kg/m3)
coordinate number ()
stress tensor (N)

Acknowledgments
This study is part of the Cluster of Excellence Unifying
Concepts in Catalysis (Unicat) (Exc 314), which is coordinated
by the Technische Universitt Berlin. The authors would like to

G.D. Wehinger et al. / Chemical Engineering Science 122 (2015) 197209

thank the Deutsche Forschungsgemeinschaft DFG within the


framework of the German Initiative of Excellence for nancial
support.
References
Bai, H., Theuerkauf, J., Gillis, P.A., Witt, P.M., 2009. A coupled DEM and CFD
simulation of ow eld and pressure drop in xed bed reactor with randomly
packed catalyst particles. Ind. Eng. Chem. Res. 48, 40604074.
Balat, H., z, C., 2007. Technical and economic aspects of carbon capture and
storage, a review. Energy Explor. Exploit. 25.
Bauer, M., Adler, R., 2003. Heat transfer in catalytic xed-bed reactors with a gas
ow, characterized through low tube/particle diameter ratios. Chem. Eng.
Technol. 26, 545549.
Behnam, M., Dixon, A.G., Nijemeisland, M., Stitt, E.H., 2013. A new approach to xed
bed radial heat transfer modeling using velocity elds from computational uid
dynamics simulations. Ind. Eng. Chem. Res. NASCRE 3.
Bey, O., Eigenberger, G., 1997. Fluid ow through catalyst lled tubes. Chem. Eng.
Sci. 52, 13651376.
Bosco, M., Vogel, F., 2006. Optically accessible channel reactor for the kinetic
investigation of hydrocarbon reforming reactions. Catal. Today 116, 348353.
Bradford, M.C.J., Vannice, M.A., 1999. CO2 reforming of CH4. Catal. Rev. 41, 142.
CD-adapco, 2014. STAR-CCM 9.02. www.cd-adapco.com.
Centi, G., Perathoner, S., 2009. Opportunities and prospects in the chemical
recycling of carbon dioxide to fuels. Catal. Today 148, 191205.
Chen, D., Ldeng, R., Anundsks, A., Olsvik, O., Holmen, A., 2001. Deactivation
during carbon dioxide reforming of methane over Ni catalyst: microkinetic
analysis. Chem. Eng. Sci. 56, 13711379.
Deutschmann, O., 2008. Computational uid dynamics simulation of catalytic
reactors In: Handbook of Heterogeneous Catalysis. Wiley-VCH Verlag GmbH
& Co. KGaA, Weinheim, pp. 18111828 (Chapter 6.6).
Dhole, S., Chhabra, R., Eswaran, V., 2006. A numerical study on the forced
convection heat transfer from an isothermal and isoux sphere in the steady
symmetric ow regime. Int. J. Heat Mass Transf. 49, 984994.
Dixon, A.G., 1997. Heat transfer in xed beds at very low ( o 4) tube-to-particle
diameter ratio. Ind. Eng. Chem. Res. 36, 30533064.
Dixon, A.G., Boudreau, J., Rocheleau, A., Troupel, A., Taskin, M.E., Nijemeisland, M.,
Stitt, E.H., 2012. Flow, transport, and reaction interactions in shaped cylindrical
particles for steam methane reforming. Ind. Eng. Chem. Res. 51, 1583915854.
Dixon, A.G., Nijemeisland, M., 2001. CFD as a design tool for xed-bed reactors. Ind.
Eng. Chem. Res. 40, 52465254.
Dixon, A.G., Nijemeisland, M., Stitt, E.H., 2013. Systematic mesh development for 3D
CFD simulation of xed beds: contact points study. Comput. Chem. Eng. 48,
135153.
Dry, M.E., 2002. The FischerTropsch process: 19502000. Catal. Today 71,
227241.
Dudukovic, M.P., 2009. Frontiers in reactor engineering. Science 325, 698701.
Eigenberger, G., 2008. Catalytic xed-bed reactors. In: Handbook of Heterogeneous
Catalysis, Wiley-VCH Verlag GmbH & Co, KGaA, Weinheim, pp. 20752106
(Chapter 10.1).
Eisfeld, B., Schnitzlein, K., 2001. The inuence of conning walls on the pressure
drop in packed beds. Chem. Eng. Sci. 56, 43214329.
Eppinger, T., Seidler, K., Kraume, M., 2011. DEM-CFD simulations of xed bed
reactors with small tube to particle diameter ratios. Chem. Eng. J. 166, 324331.
Ergun, S., 1952. Fluid ow through packed columns. Chem. Eng. Prog. 48, 8994.
Freund, H., Bauer, J., Zeiser, T., Emig, G., 2005. Detailed simulation of transport
processes in xed-beds. Ind. Eng. Chem. Res. 44, 64236434.
Geske, M., Korup, O., Horn, R., 2013. Resolving kinetics and dynamics of a catalytic
reaction inside a xed bed reactor by combined kinetic and spectroscopic
proling. Catal. Sci. Technol. 3, 169175.
Giese, M., Rottschafer, K., Vortmeyer, D., 1998. Measured and modeled supercial
ow proles in packed beds with liquid ow. AIChE J. 44, 484490.
Ginsburg, J.M., Pia, J., El Solh, T., de Lasa, H.I., 2005. Coke formation over a nickel
catalyst under methane dry reforming conditions: thermodynamic and kinetic
models. Ind. Eng. Chem. Res. 44, 48464854.
Grf, I., Rhl, A.K., Kraushaar-Czarnetzki, B., 2014. Experimental study of heat
transport in catalytic sponge packings by monitoring spatial temperature
proles in a cooled-wall reactor. Chem. Eng. J. 244, 234242.
Guardo, A., Coussirat, M., Larrayoz, M., Recasens, F., Egusquiza, E., 2005. Inuence of
the turbulence model in CFD modeling of wall-to-uid heat transfer in packed
beds. Chem. Eng. Sci. 60, 17331742.
Guo, J., Lou, H., Zhao, H., Chai, D., Zheng, X., 2004. Dry reforming of methane over
nickel catalysts supported on magnesium aluminate spinels. Appl. Catal. A:
Gen. 273, 7582.
Guo, J., Lou, H., Zheng, X., 2007. The deposition of coke from methane on a Ni/
MgAl2O4 catalyst. Carbon 45, 13141321.
Hartmann, D., Tank, A.K., Rusticucci, M., Alexander, L., Brnnimann, S., Charabi, Y.,
Dentener, F., Dlugokencky, E., Easterling, D., Kaplan, A., Soden, B., Thorne, P.,
Wild, M., Zhai, P., 2013. Observations: atmosphere and surface. In: IPCC Fifth
Assessment Report: Climate Change 2013 (AR5): The Physical Science Basis.
Contribution of Working Group I to the Fifth Assessment Report of the
Intergovernmental Panel on Climate Change. Cambridge University Press,
Cambridge, United Kingdom, New York, NY, USA, pp. 159254.

209

Hartmann, M., Maier, L., Minh, H., Deutschmann, O., 2010. Catalytic partial
oxidation of iso-octane over rhodium catalysts. Combust. Flame: Exp. Model.
Simul. Study 157, 17711782.
Hickman, D.A., Schmidt, L.D., 1993. Steps in CH4 oxidation on Pt and Rh surfaces:
high-temperature reactor simulations. AIChE J. 39, 11641177.
Horn, R., Korup, O., Geske, M., Zavyalova, U., Oprea, I., Schlgl, R., 2010. Reactor for
in situ measurements of spatially resolved kinetic data in heterogeneous
catalysis. Rev. Sci. Instrum. 81, 064102.
Hunt, A.J., Sin, E.H.K., Marriott, R., Clark, J.H., 2010. Generation, capture, and
utilization of industrial carbon dioxide. ChemSusChem 3, 306322.
Kahle, L.C.S., Roussire, T., Maier, L., Herrera Delgado, K., Wasserschaff, G., Schunk, S.
A., Deutschmann, O., 2013. Methane dry reforming at high temperature and
elevated pressure: impact of gas-phase reactions. Ind. Eng. Chem. Res. 52,
1192011930.
Kee, R.J., Colin, M.E., Glarborg, P., 2003. Chemically Reacting Flow, Theory and
Practice. Wiley, Hoboken, NJ.
de Klerk, A., 2003. Voidage variation in packed beds at small column to particle
diameter ratio. AIChE J. 49, 20222029.
Korup, O., Mavlyankariev, S., Geske, M., Goldsmith, C.F., Horn, R., 2011. Measurement and analysis of spatial reactor proles in high temperature catalysis
research. Chem. Eng. Process.: Process Intensif. 50, 9981009.
Kuroki, M., Ookawara, S., Ogawa, K., 2009. A high-delity CFD model of methane
steam reforming in a packed bed reactor. J. Chem. Eng. Jpn. 42, s73s78.
Leva, M., Weintraub, M., Grummer, M., Pollchik, M., Storch, H.H., 1951. Fluid ow
through packed and uidized systems. US Bur. Mines Bull. 504.
Maestri, M., Vlachos, D.G., Beretta, A., Groppi, G., Tronconi, E., 2008. Steam and dry
reforming of methane on Rh: microkinetic analysis and hierarchy of kinetic
models. J. Catal. 259, 211222.
Maestri, M., Vlachos, D.G., Beretta, A., Groppi, G., Tronconi, E., 2009. A C1
microkinetic model for methane conversion to syngas on Rh/Al2O3. AIChE J.
55, 9931008.
McGuire, N.E., Sullivan, N.P., Deutschmann, O., Zhu, H., Kee, R.J., 2011. Dry reforming
of methane in a stagnation-ow reactor using Rh supported on strontiumsubstituted hexaaluminate. Appl. Catal. A: Gen. 394, 257265.
Mhadeshwar, A.B., Vlachos, D.G., 2005. Hierarchical multiscale mechanism development for methane partial oxidation and reforming and for thermal decomposition of oxygenates on Rh. J. Phys. Chem. B 109, 1681916835.
Mikkelsen, M., Jorgensen, M., Krebs, F.C., 2010. The teraton challenge. A review of
xation and transformation of carbon dioxide. Energy Environ. Sci. 3, 4381.
Mueller, G.E., 1992. Radial void fraction distributions in randomly packed xed
beds of uniformly sized spheres in cylindrical containers. Powder Technol. 72,
269275.
Ookawara, S., Kuroki, M., Street, D., Ogawa, K., 2007. High-delity DEM-CFD
modeling of packed bed reactors for process intensication. In: Proceedings
of European Congress of Chemical Engineering (ECCE-6), Copenhagen.
Papadopoulou, C., Matralis, H., Verykios, X., 2012. Utilization of biogas as a
renewable carbon source: dry reforming of methane. In: Catalysis for Alternative Energy Generation. Springer, New York, pp. 57127.
Reichelt, W., 1972. Zur Berechnung des Druckverlustes einphasig durchstrmter
Kugel- und Zylinderschttungen. Chem. Ingen. Tech. 44, 10681071.
Ren, Z., Pope, S.B., 2008. Second-order splitting schemes for a class of reactive
systems. J. Comput. Phys. 227, 81658176.
Rostrup-Nielsen, J., Hansen, J., 1993. CO2-reforming of methane over transition
metals. J. Catal. 144, 3849.
Salciccioli, M., Stamatakis, M., Caratzoulas, S., Vlachos, D., 2011. A review of
multiscale modeling of metal-catalyzed reactions: mechanism development
for complexity and emergent behavior. Chem. Eng. Sci. 66, 43194355.
Schwiedernoch, R., Tischer, S., Correa, C., Deutschmann, O., 2003. Experimental and
numerical study on the transient behavior of partial oxidation of methane in a
catalytic monolith. Chem. Eng. Sci. 58, 633642.
Shih, T.H., Liou, W.W., Shabbir, A., Yang, Z., Zhu, J., 1995. A new k- eddy viscosity
model for high Reynolds number turbulent ows. Comput. Fluids 24, 227238.
Taskin, M.E., Troupel, A., Dixon, A.G., Nijemeisland, M., Stitt, E.H., 2010. Flow,
transport, and reaction interactions for cylindrical particles with strongly
endothermic reactions. Ind. Eng. Chem. Res. 49, 90269037.
Torniainen, P., Chu, X., Schmidt, L., 1994. Comparison of monolith-supported metals
for the direct oxidation of methane to syngas. J. Catal. 146, 110.
Vortmeyer, D., Winter, R., 1984. On the validity limits of packed bed reactor
continuum models with respect to tube to particle diameter ratio. Chem. Eng.
Sci. 39, 14301432.
Wang, S., Lu, G.Q.M., Millar, G.J., 1996. Carbon dioxide reforming of methane to
produce synthesis gas over metal-supported catalysts: state of the art. Energy
Fuels 10, 896904.
Wehinger, G.D., Eppinger, T., Kraume, M., 2014. Fluidic effects on kinetic parameter
estimation in lab-scale catalysis testinga critical evaluation based on computational uid dynamics. Chem. Eng. Sci. 111, 220230.
Wolfshtein, M., 1969. The velocity and temperature distribution in one-dimensional
ow with turbulence augmentation and pressure gradient. Int. J. Heat Mass
Transf. 12, 301318.
Zikowska, I., Zikowski, D., 1988. Fluid ow inside packed beds. Chem. Eng.
Process.: Process Intensif. 23, 137164.
Zobel, N., Eppinger, T., Behrendt, F., Kraume, M., 2012. Inuence of the wall
structure on the void fraction distribution in packed beds. Chem. Eng. Sci. 71,
212219.

Você também pode gostar