Journal Pre-proof Heat transfer characteristics analysis and optimization of the Micro Packed Bed Bionic Reactor Wei-liao Liu Xi Li You-wei Cheng Li-jun Wang
PII:
S0255-2701(19)30402-7
DOI:
https://doi.org/doi:10.1016/j.cep.2019.107644
Reference:
CEP 107644
To appear in:
Chemical Engineering and Processing
Received Date:
10 April 2019
Revised Date:
6 August 2019
Accepted Date:
21 August 2019
Please cite this article as: Wei-liao Liu, Xi Li, You-wei Cheng, Li-jun Wang, Heat transfer characteristics analysis and optimization of the Micro Packed Bed Bionic Reactor, (2019), doi: https://doi.org/10.1016/j.cep.2019.107644
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2019 Published by Elsevier.
f oo pr
Figure 1: (a) Schematic diagram of mammal circulatory system, (b) Simplified schematic
Jo u
rn
al
Pr
e-
diagram of the hierarchical channel system, (c) Elementary level of the FBBR.
Figure 2: Sectional view of the prototype elementary structure of the MPBBR.
1
Page 1 of 36
f oo pr
e-
Figure 3: (a) Sectional view of the MPBBR elementary unit, 2D geometries and boundary
Jo u
rn
al
Pr
conditions of x − y plane (b) and x − z plane (c).
Figure 4: Heat transfer interactions between (a) radial conduction and convection in the packed bed, (b) axial convection and radial conduction in the input channel, (c) axial conduction and reaction heat in the packed bed.
2
Page 2 of 36
f oo pr
Figure 5: Influence of wall thickness on the axial temperature distribution in the input channel
Jo u
rn
al
Pr
e-
and catalyst bed.
Figure 6: Influence of channel width on the axial temperature distribution in the input channel and catalyst bed.
3
Page 3 of 36
f oo pr e-
Jo u
rn
al
Pr
Figure 7: Sectional view of the MPBBR structure with back-turning input channel.
Figure 8: Heat transfer behavior and temperature profile in (a) the normal single input channel, (b) the back-turning input channel (2h = 3mm, N = 4, ww = 0.6mm, L = 10cm).
4
Page 4 of 36
f oo
Figure 9: Axial distribution of channel temperature with and without back-turning channel
al
Pr
e-
pr
(2h = 3mm, N = 4, ww = 0.6mm, L = 10cm).
Figure 10: Relation between maximal bed temperature difference and bed length with and
Jo u
rn
without back-turning channel (2h = 3mm, N = 4, ww = 0.6mm).
Figure 11: Comparisons of axial temperature distribution between other monolithic reactor investigations and the MPBBR (single input channel).
5
Page 5 of 36
f oo
rn
al
Pr
e-
pr
Figure 12: Schematic and geometric parameters of the HRMC (unit:mm).
Jo u
Figure 13: Comparisons of axial temperature distribution between the HRMC and CMC.
Figure 14: Top view of the MPBBR structure with back-turning input channel.
6
Page 6 of 36
f oo pr e-
Figure 15: Relation between the maximal temperature difference and bed length for different
Jo u
rn
al
Pr
channel density (y0,CH4 = 1%).
Figure 16: Relation between the maximal temperature difference and bed length for different channel density (y0,CH4 = 0.5%).
7
Page 7 of 36
f oo pr ePr al rn
Figure 17: Temperature profile for the whole elementary unit, framework domain, and catalyst
Jo u
bed domains under optimized conditions (y0,CH4 = 1%).
8
Page 8 of 36
ur Jo
Page 9 of 36
Introduction of the Micro Packed Bed Bionic Reactor (MPBBR) Mathematical modeling of MPBBR elementary structure unit Analysis of the heat transfer behaviors and auto-thermal function MPBBR design for the catalytic combustion of lean methane
Jo u
rn
al
Pr
e-
pr
oo
f
Optimization of structure and geometric parameters
Page 10 of 36
Heat transfer characteristics analysis and optimization of the Micro Packed Bed Bionic Reactor Wei-liao Liu, Xi Li∗, You-wei Cheng, Li-jun Wang
oo
f
College of Chemical and Biological Engineering, Zhejiang University, 38 Zheda road, Hangzhou 310027, People’s Republic of China
Abstract
pr
The Fixed Bed Bionic Reactor (FBBR) is a novel conceptual reactor using hierarchical flow channel networks to reduce the pressure drop in fixed beds so
e-
that fine catalysts can be used to enhance the catalyst effectiveness. The Micro Packed Bed Bionic Reactor (MPBBR) is a representative implementation of the FBBR designed for fast catalytic reactions with strong heat effect. In this
Pr
work, the heat transfer behaviors of the MPBBR are qualitatively analyzed. By a typical application case, the catalytic combustion of lean methane, the influences of channel structure and geometric parameters on the heat transfer
al
performance are quantitatively analyzed. An optimal MPBBR for such highly exothermic reaction uses porous silicon carbide ceramic monoliths as honeycomb
rn
framework. The single input channel is modified by add a back-turning channel. With the optimized channel structure and geometric parameters, the maximal
Jo u
temperature difference in the catalyst bed is maintained under 45o C, and the unit catalyst mass and unit reactor volume handling capacities are doubled and tripled, respectively. The MPBBR is proved to be a flexible and economical device with high performance heat transfer. Keywords:
Fixed bed bionic reactor, Heat transfer in porous channel, Auto-thermal function, Reactor optimization, Catalytic combustion of lean methane ∗ Corresponding
author Email address:
[email protected] (Xi Li)
Preprint submitted to Chemical Engineering & Processing: Process IntensificationAugust 6, 2019
Page 11 of 36
Nomenclature
2
I~
direction vector
3
~u
velocity vector, m/s
4
B
permeability, m2
5
C
molar concentration, mol/m3
6
Cp
heat capacity at constant pressure, J/kg · K
7
D
diffusivity, m2 /s
8
dp
catalyst particle diameter, mm
9
h
half channel width, mm
10
K
thermal conductivity, W/m · K
11
L
bed length, cm
12
N
column number of catalyst bed
13
p
pressure, P a
14
QR
reaction heat, W/m3
15
R
reaction rate, mol/m3 · s
16
u
flow velocity in x-direction, m/s
17
v
flow velocity in y-direction, m/s
18
w
thickness, mm
19
Greek symbols
20
porosity
21
µ
dynamic viscosity, P a · s
Jo u
rn
al
Pr
e-
pr
oo
f
1
2
Page 12 of 36
0
feed
25
b
catalyst bed
26
e
effective coefficient
27
g
gaseous reactant
28
L
longitudinal direction
29
s
solid material
30
T
transverse direction
31
w
permeable wall
32
1. Introduction
oo
24
pr
Subscript
e-
23
Pr
ρ
f
density, kg/m3
22
The Fixed Bed Bionic Reactor (FBBR) is a novel conceptual reactor. The
34
core idea of the FBBR is the use of hierarchical input and output channel net-
35
works to reduce the pressure drop in packed beds so that fine catalysts can be
36
used to enhance the catalyst effectiveness [1]. The FBBR structure is inspired
37
from the mammal circulatory system, which consists of artery and vein channel
38
networks and capillaries in Fig. 1(a). Fig. 1(b) is the simplified schematic di-
39
agram demonstrating the hierarchical channel system from aorta to vena cava.
40
The capillaries only exist between the lowest level of arteries and veins. The
41
FBBR adopts the distributed flow pattern which is realized by proper arrange-
42
ment of different levels of flow channels. The elementary structure level is shown
43
in Fig. 1(c). As the arrows indicate, the distributed reactant fluids flow verti-
44
cally into the close-ended input channels, penetrate the adjacent catalyst beds,
45
and leave through the nearest output channels. The input and output channels
46
are alternatively configured. The catalyst particles are packed in the space area
Jo u
rn
al
33
3
Page 13 of 36
47
between adjacent channels. In this way the catalyst bed can be segmented to
48
rather thin layers and the infiltration velocity is greatly reduced due to large
49
contact surface between the channels and the beds, resulting in low flow resis-
50
tance and the applicability of fine catalysts. Based on different reaction conditions and production scales, there are sev-
52
eral kinds of implementations of the FBBR. The MPBBR is a representative
53
implementation and its elementary structure is formed by honeycomb frame-
54
works. Fig. 2 shows a sectional view of the prototype structure of the MPBBR.
55
One end of a column is sealed to form input or output channel. The catalysts
56
are packed and sealed in several columns between a pair of input and output
57
channels to form the catalyst bed. The manufacture of the MPBBR is simple,
58
since the honeycomb framework can be realized by wall-flow monoliths. Due
59
to the small scale and compact structure of flow channels and catalyst beds,
60
the heat transfer resistance is fundamentally reduced in the MPBBR. There-
61
fore, with proper design to enhance the heat transfer performance, it may be
62
an attractive option for fast catalytic reactions with strong heat effect.
Pr
e-
pr
oo
f
51
Several investigations were made upon heat transfer enhancement in uncon-
64
ventional packed bed reactors. Vervloet et al. [2] compared several structured
65
packing elements for tubular fixed bed reactors and suggested packed closed
66
cross flow structure was good in heat transfer performance and catalyst hold-
67
up. Visconti et al. [3] used highly conductive metal-foam to enhance the heat
68
conduction in the packed bed. Busse et al. [4] investigated the periodic open
Jo u
rn
al
63
69
cellular structures and demonstrated their beneficial properties like high specific
70
surface area, low pressure drop, and adjustable heat transfer properties. How-
71
ever, the MPBBR has the small scale structure of flow channels and packed beds,
72
of which the widths and thicknesses are in millimeter level. Thus it is difficult
73
and uneconomical to embed other heat transfer structures like conductive foams
74
or to use external heat exchange units like tube bundles, which means the heat
75
transfer enhancement in the MPBBR will be merely based on the optimization
76
of the channel structure and geometric parameters. Therefore, it is important
77
to study the heat transfer behaviors in the MPBBR and to quantitatively an4
Page 14 of 36
78
alyze the influence of the channel structure and the geometric parameters on
79
the heat transfer performance, so that with proper choices of the structure and
80
geometric parameters, the temperature gradient in the beds can be limited to a
81
tolerant range. The uncoupled flow and heat transfer research works can trace back to Prins
83
et al. [5] for flow and heat transfer between parallel plates with constant wall
84
temperature. Cess and Shaffer [6] consider the same problem with prescribed
85
wall heat flux. Terrill [7] further considered flow and transfer between porous
86
plates with small suction Reynolds numbers and Yeroshenko et al. [8] extended
87
the results to large P eclet numbers with both suction and injection. Recent
88
analytical studies focused on the heat transfer between porous parallel plates
89
filled with a saturated porous medium and with asymmetric uniform heat flux
90
[9, 10]. However, of all the previous works focused on heat transfer between
91
parallel plates with given wall temperature or heat flux, none considered flow
92
and heat transfer in the channels coupled with reaction heat. Rebrov et al. [11]
93
and Haber et al. [12] investigated the heat transfer by numerical methods in the
94
microstructured reactor, where the catalysts are coated on channel walls. The
95
investigation upon heat transfer in a system coupled by porous parallel plates
96
and adjacent packed bed with unprescribed wall heat flux boundary conditions
97
is not reported in literatures. The transport phenomenon in this coupled system
98
is complicated and nonlinear, so in this work, numerical simulations are used
99
to quantitatively estimate the temperature distribution and to optimize the
Jo u
rn
al
Pr
e-
pr
oo
f
82
100
channel structure and geometric parameters.
101
2. Model description and analysis
102
2.1. Mathematical models of the MPBBR
103
The 3D geometric model of an elementary unit is shown in Fig. 3(a) and
104
the types of the boundary conditions are shown in the sectional views (Fig. 3
105
(b) and (c)). The 3D model is divided into 3 kinds of domains: the free flow
106
domains of the flow channels; the porous media domains of the packed beds;
5
Page 15 of 36
the one single porous medium domain of the framework that contains all the
108
permeable walls. The major assumption in the models are listed as follows:
109
(1) The fluid is ideal gas and is incompressible.
110
(2) Pseudo-homogeneous model is used.
111
(3) The catalyst beds and the framework are considered as individual uniform
112
porous media with constant porosity, permeability, and thermal conductivity.
113
(4) The operation is adiabatic.
114
2.1.1. Momentum transfer model
oo
f
107
The flow in the free channels is described by the stationary and incompress-
116
ible Navier-Stokes equations, while the flow in porous media is described by
117
the Brinkman equations with Forchheimer correction [13]. The coupled Stokes-
118
Darcy-Brinkman model was used for coupled free and porous media flow [14, 15].
119
The earlier studies of this coupled model by Beavers and Joseph [16] exhibited
120
the existence of a slip velocity at the fluid/porous interface. An interface bound-
121
ary condition was given by continuity to satisfy the slip velocity and shear stress
122
at the interface [17]. However, Schmitz and Prat [18] showed that the slip effect
123
can be neglected for low porosity surface like membrane. In the MPBBR, the
124
permeable wall is also a kind of low porosity surface ( < 30%), so the slip effect
125
is also neglected at the channel/wall interface. The momentum transfer model
e-
Pr
al
rn
is expressed as eq. (1) & (2). Continuity eq. : O · ~u = 0 F ree channel : ρ(~u · O)~u = O · [−pI~ + µ(O~u + (O~u)T )]
Jo u
126
pr
115
Catalyst bed : ρ2 (~u · O)~u = O · [−pI~ + µb (O~u + (O~u)T )] − Bµb ~u − βF |~u|~u b P ermeable wall : ρ (~u · O)~u = O · [−pI~ + µ (O~u + (O~u)T )] − µ ~u − β |~u|~u F 2w w Bw (1)
127
Boundary conditions: Inlet : ~u = u0 I~x Outlet : p = ambient pressure Channel/wall interf ace : ~u = (0, Bw Op · I~ , 0) y
(2)
µ
6
Page 16 of 36
128
Here the inlet velocity is assumed to be perpendicular to the inlet surface and
129
backflow is suppressed at the outlet surface. B is the permeability of the porous
130
media estimated by eq. (3) [19]. B=
d2p 3 150(1 − )2
(3)
βF |~u|~u is the Forchheimer correction for turbulent drag contributions. βF is the
132
non-Darcy flow coefficient estimated by an empirical correlation eq. (4) [20].
oo
f
131
βF = 5.5 × 10−12 B −1.47 −0.53 ρ 2.1.2. Mass transfer model
pr
133
(4)
The mass transfer is described by the convection and diffusion model. In
135
the porous media, the effective diffusivity and mass dispersion are considered. The mass transfer model is expressed as eq. (5) & (6) F ree channel : O · (−Di OCi ) + ~u · OCi = 0 Catalyst bed : O · (−(D + D )OC ) + ~u · OC = R e,i D,i i i i P ermeable wall : O · (−(D + D )OC ) + ~u · OC = 0 e,i D,i i i i = CH , O , H O, CO , N 4 2 2 2 2 Inlet : C = C i 0,i Outlet : D OC = 0 i
(6)
i
The calculation correlations of different diffusivities are listed in Table 1. Here
Jo u
138
(5)
Boundary conditions:
rn
137
al
Pr
136
e-
134
139
P eM is the mass transfer Peclet Number P eM = ul/D and Sc is the Schmidt
140
Number Sc = µ/ρD, where l is the characteristic length. For practical applica-
141
tion cases of the MPBBR, 10 < P eM < 500 and 1 < Sc < 200.
142
2.1.3. Energy transfer model
143
The energy transfer is described by the convection and diffusion model. The
144
third boundary conditions are given for the inlet surface and the channel/wall
145
interface. In the porous media, the thermal dispersion is considered. The energy
7
Page 17 of 36
Table 1: Calculation correlations for diffusivities (unit:m2 /s).
Diffusivity in mixture [19] Di =
P P 1−yi ( Nj )/Ni (yj −yi Nj /Ni )/Dij
Binary diffusivity [19] Dij =
0.01T 1.75 (1/Mi +1/Mj )0.5 P P p[( Vi )1/3 +( Vj )1/3 ]2
De = τ Di
oo
f
Effective diffusivity [19]
Longitudinal (y-direction) dispersion coefficient [21] DL = P eM Di /(25Sc1.14 /P eM + 0.5)
pr
Transverse (x,z-direction) dispersion coefficient [21]
transfer model is expressed as eq. (7) & (8) ρCp ~u · OT − O · (KOT ) = Q F low channels : Q = 0, K = K , g Catalyst beds : Q = QR , K = Ke,b + Kdisp P ermeable wall : Q = 0, K = K + K e,w disp Boundary conditions: Inlet : Kg OT · I~x = ρCp u0 (T − T0 ) Outlet : Kg OT = 0 Channel/wall interf ace : K OT · I~ = h ∆T e,w y w
(8)
Jo u
rn
147
(7)
al
Pr
146
e-
DT = (1 + (2.7 × 10−5 Sc + 12/P eM )−1 )Di
148
The thermal conductivity of gaseous flow Kg is calculated by molar average.
149
The effective thermal conductivities of porous media are estimated by Hadley’s
150
correlations eq. (9) & (10) [22]. Ke 0.8 + (Ks /Kg )(1 − 0.8) 2(1 − )(Ks /Kg )2 + (1 + 2)(Ks /Kg ) = (1−α0 ) +α0 Kg 1 − 0.2(Ks /Kg − 1) (2 + )(Ks /Kg ) + 1 − (9)
151
where log α0 = −1.084 − 6.778( − 0.298)
0.298 ≤ ≤ 0.580
(10)
8
Page 18 of 36
152
The thermal dispersion coefficients are estimated by Saffman’s correlations eq. (11)
153
& (12) [23]. DL P eH P eH 17 = [ln(122P eH ) − − ] Kg 6 200 12
(11)
DT 3P eH P e2H = + Kg 16 1000
(12)
154
Here P eH is the heat transfer Peclet Number P eH = ulρCp /K, where l is the
156
characteristic length. For practical application cases of the MPBBR, 0.01 <
157
P eH < 0.5.
oo
f
155
hw is the effective heat transfer coefficient at channel/wall interface, which
159
is determined by the channel Nusselt number N u. Terrill and Yeroshenko [7, 8]
160
studied the heat transfer in flow between parallel porous plates and gave a
161
correlation for N u as eq. (13). hw is then estimated by eq. (14), where δ is the
162
thickness of the heat transfer layer.
Pr
e-
pr
158
N u = 3.77 + P eH + 0.087P e2H + 0.01Rew 163
164
(13) (14)
al
hw = N uKg /δ
(P eH < 3)
2.2. Model credibility discussion
The assumptions listed in section 2.1 are all common assumptions. No
166
practical application case for the Bionic Reactor is under very high pressure
167
or low temperature so the ideal gas assumption should be valid. The pseudo-
Jo u
rn
165
168
homogeneous model and the uniform porous media assumption are commonly
169
used in the simulation studies of gas-solid catalytic reactions in packed beds
170
[24, 25, 26, 27]. The MPBBR is an adiabatic reactor and can be considered as
171
an integration of many individual elementary units, so it is reasonable to assume
172
adiabatic operation in the elementary unit.
173
The coupled Stokes-Darcy-Brinkman model with negligible interface bound-
174
ary condition was commonly used for membrane processes [28, 29]. Damak [30]
175
proved that this model was accurate enough for a membrane process with chan-
176
nel Reynolds number between 300 and 1000 and infiltration Reynolds number 9
Page 19 of 36
between 0.1 and 0.3. The flow in the MPBBR may cover a larger range of
178
Reynolds number, so the Forchheimer correction for turbulent drag contribution
179
is also considered. Thus the model is still valid when the infiltration deviates
180
from Darcy flow. The convection and diffusion models of mass and energy trans-
181
fer for coupled domains with different physical properties are commonly used
182
in microchannel and membrane reactors [31, 32, 33, 34].The basic behaviors of
183
the mass and energy transfer of the MPBBR are similar to the membrane and
184
microchannel reactors and the effective diffusivities and thermal conductivities
185
are estimated respectively for free flow and porous domains. The mass and
186
thermal dispersions are also considered for the porous domains. Therefore, the
187
current model can be considered of enough credibility.
188
2.3. Flow distribution criterion
e-
pr
oo
f
177
The flow distribution characteristics was studied analytically in a 2D model
190
in the authors’ earlier publication [1]. A dimensionless parameter Pr (Eq. (15))
191
was identified to be the key factor that determining the flow distribution unifor-
192
mity. Another parameter, the infiltration Reynolds number Rew , was demon-
193
strated to play a second role. Relations between Pr , Rew and the relative
194
maximal infiltration velocity difference σmax were obtained from the analytical
195
solution of the model, and this holds for the flow distribution in the MPBBR.
196
However, in the previous work, the influence of the permeable walls were ne-
197
glected. In this work, the 3D geometric model of an elementary unit in Fig. 3(a)
rn
al
Pr
189
shows that a catalyst bed consists of several columns of channels, so the influ-
199
ence of the permeable wall should be considered. Therefore, some corresponding
200
modifications need to be made in the expression of Pr .
Jo u 198
Pr =
u0 ρBb L µhwb
(15)
hvw ρ µ
(16)
201
Rew = 202
Assuming that the catalyst consists of N columns of channel, then the
203
total pressure difference between the input and output channel is estimated by
10
Page 20 of 36
204
eq. (17). ∆p = N
205
2hvw µ ww vw µ + (N + 1) Bb Bw
(17)
Thus the expression of Pr number of the MPBBR is modified as Eq. (18). Pr =
u0 ρBbw L 2h2 µN
(18)
2hN ww (N + 1) −1 + ) Bb Bw
(19)
oo
Bbw = (
f
206
207
where the permeabilities of the packed bed and porous wall are esitimated by
208
Eq. (3), and Bbw can be considered as the average unit permeability. The modified Pr number now includes all the independent geometric pa-
210
rameters and physical properties. Since the Pr number is the key parameter
211
that determines the flow distribution uniformity, it should always be an indis-
212
pensable criterion of the optimal design of the MPBBR.
213
2.4. Heat transfer behaviors in the MPBBR
Pr
e-
pr
209
Although analytic solutions are unavailable for the coupled mathematical
215
model of the MPBBR, several heat transfer interactions can be demonstrated
216
by Fig. 4 to help comprehending the heat transfer behaviors.
al
214
In a common case where the reaction is exothermic and the feed fluid is
218
cold, the radial conduction and convection are in opposite directions, as shown
219
in Fig. 4(a). In a conventional fixed bed reactor, the large size catalyst bed
220
with high infiltration velocity will result in large heat conduction resistance and
Jo u
rn
217
221
the domination of the convection of the cold inflow. Thus the feed fluid must
222
be sufficiently preheated to avoid extinction. Whereas, in the MPBBR, the bed
223
size and infiltration velocity can be one or two orders of magnitude lower for
224
the heat conduction to become dominant, resulting in a smaller temperature
225
gradient in the catalyst bed. Thus in the MPBBR, cold feed can be directly
226
used without preheating which is essential for the realization of the auto-thermal
227
function, since the reaction heat can be neutralized by the sensible heat increase
228
of the cold feed. However, the axial temperature difference along the input
229
channel will be inevitable because of the interaction between axial convection 11
Page 21 of 36
and radial conduction in Fig. 4(b), which is denoted by the channel Peclet
231
Number P e = ρCp uh/K. The characteristic length h here refers to the input
232
channel width, meaning that a wider input channel will be in favor of limiting
233
the radial conduction so that the axial difference in the channel temperature
234
can be reduced. By contrast, the interaction between the axial conduction and
235
reaction heat in the packed bed is more intuitional in Fig. 4(c). A higher bed
236
heat conductivity or a shorter bed length can generally result in a more uniform
237
bed temperature distribution.
oo
f
230
The aforementioned heat transfer behaviors imply that thinner and shorter
239
catalyst bed and wider input channel can result in perfect uniformity of tem-
240
perature distribution. However, there are other indispensable restrictions to
241
follow, like the Pr number, the pressure drop, the catalyst packing fraction, the
242
handling capacity, and the manufacture feasibility as well. In section 3, an ap-
243
plication case will be introduced to quantitatively estimate the influence of the
244
channel structure and geometric parameters on the temperature distribution
245
uniformity as well as to demonstrate the optimization of the MPBBR.
246
3. Analysis and optimization of application case
al
Pr
e-
pr
238
The catalytic combustion of lean methane (concentration of 0.5 – 1v%) is a
248
representative fast catalytic highly exothermic reaction. The present industrial
249
processes use the catalytic flow reversal reactors (CFRR) loaded with cordierite
250
monoliths [35, 36]. The Reverse-Flow Operation for the catalytic combustion
251
requires a vast amount of room for thermal storage media, resulting in a catalyst
252
packing fraction of lower than 25%. Since the space velocity is as high as
253
36000h−1 , to maintain a low pressure drop, large catalyst particles have to be
254
used. In the CFRR, Raschig rings with a characteristic length of 7.5mm are
255
used and its catalyst effectiveness is only about 42.1%. In this situation, to
256
improve the handling capacity and in the meantime satisfy the heat transfer
257
requirement, the MPBBR appears to be an ideal substitution.
Jo u
rn
247
12
Page 22 of 36
258
3.1. Advantages and operation conditions for the MPBBR
259
Since the pressure drop is fundamentally reduced in the MPBBR, under
260
the same level of pressure drop as the CFRR, fine catalysts with much higher
261
effectiveness can be used. To eliminate the resistance of the intra-particle dif-
262
fusion, 0.18mm fine catalysts are used and its catalyst effectiveness can reach
263
over 90%. The adiabatic temperature rise of this reaction is about 340o C for 1v%
265
concentration of methane, so it is feasible to use cold feed without preheating
266
and take advantage of the auto-thermal function of the MPBBR. Thus the
267
thermal storage media are no longer needed and the catalyst packing fraction
268
can be increased. In recent investigations, high performance catalysts have been
269
found for low temperature catalytic combustion. The ignition temperature is
270
lower than 350o C and over 99% conversion of methane can be achieved under
271
570o C [37, 38]. Therefore, the inlet temperature is set to be 230o C for the heat
272
balance.
Pr
e-
pr
oo
f
264
In the following sections, the heat transfer characteristics will be quantita-
274
tively analyzed based on the numerical solutions to the coupled mathematical
275
model equations of the MPBBR (Eq. (1) – (14)) through Comsol M ultiphysicsT M .
276
The feed stream for the simulations contains 79v% N2 , 20v% O2 ,1v% CH4 , and
277
the inlet temperature and pressure are 230o C and 0.103M P a. The inlet velocity
278
condition varies based on different geometric parameters.
rn
al
273
3.2. Material selection for the honeycomb framework
Jo u 279
280
The cordierite ceramic is a commonly used material for monoliths. It has
281
good permeability and mechanical strength. However, for porosity of 24%,
282
the cordierite wall thermal conductivity is only 1.5W/(mK) [39, 40], which is
283
evidently not applicable for such highly exothermic reaction. For higher thermal
284
conductivity, the metal foam and metal matrix composites (MMCs) seem to be
285
a better option, since the specific thermal conductivities range from 10W/(mK)
286
to over 100W/(mK) for different metals [41]. Whereas, the defects are that the
287
performance of metal foam and MMCs can’t hold under high temperature (over 13
Page 23 of 36
400o C) and the manufacture is rather expensive. Thus the porous silicon carbide
289
ceramic tends to be a wise choice, for it is a high temperature resistant and
290
economic material with considerable thermal conductivity. According to Liu &
291
Tuan [42], the thermal conductivity of such material prepared by mixing coarse
292
SiC particles with a small amount of sodium silicate can reach 19W/(mK), even
293
though the porosity is as high as 40%.
294
3.3. Heat transfer characteristics analysis
oo
f
288
Since the inlet temperature is 120o C lower than the ignition temperature,
296
and on the other hand, the catalytic activity shows a typical decrease above
297
580o C for the P d/Cex Zr1−x O2 catalyst [37], the temperature control of the
298
MPBBR should be precise enough to avoid extinction or deactivation of cat-
299
alysts. This requires reasonable optimizations of the channel structure and
300
geometric parameters. The geometric parameters to be optimized include the
301
total catalyst bed thickness wb , the whole channel width 2h, the wall thickness
302
ww , and the bed length L.
Pr
e-
pr
295
Based on the qualitative judgement in section 2.4, thinner bed thickness
304
is in favor of better heat transfer performance and lower pressure drop, while
305
some trial simulations show that the axial temperature difference in the cata-
306
lyst bed is the main contribution to the non-uniformity of temperature, so for
307
the practical interests, the determination of wb should concern more about the
308
catalyst packing fraction and handling capacity. According to the pressure drop
rn
al
303
restriction, wb can be at most set to be 12mm to maintain the same level of
310
pressure drop as that in the CFRR. The corresponding infiltration velocity is
311
0.12m/s, which is fast enough to eliminate the resistance of the inter-particle
312
diffusion.
Jo u 309
313
The prototype of the honeycomb framework uses single column as input
314
channel (Fig. 2). Trial simulations show that the main temperature difference is
315
in the axial direction. Thus to quantitatively estimate the non-uniformity of the
316
axial temperature distribution, simulations are carried out for elementary units
317
with L = 10cm, 2h = 3, ww = 0.5, 0.6, 0.8mm, and L = 10cm, ww = 0.6, 2h = 14
Page 24 of 36
2, 3, 4mm. Fig. 5 indicates that the wall thickness has little influence on the
319
channel temperature Tc distribution, whereas the axial difference of the bed
320
temperature Tb decreases as the wall thickness increases. This is predictable
321
because the SiC framework is the main contributor of the heat transfer, and
322
thicker wall will lead to higher equivalent heat conductivity, especially in the
323
axial direction [43, 44]. Fig. 6 indicates that the axial channel temperature
324
difference is significantly affected by the channel width. As discussed in section
325
2.4, the axial heat convection become dominant in a wider flow channel, so the
326
channel temperature will be more uniform. We can also tell from the variation
327
of the Tb curves that reducing the channel temperature difference is an effective
328
way to have a more uniform bed temperature distribution. However, merely
329
increasing the width of channels and walls will not be acceptable for catalyst
330
packing fraction or reactor manufacture feasibility.
e-
pr
oo
f
318
Since a large temperature rise will always exist in the first pass through
332
the input channel, the input channel can be modified by adding a back-turning
333
channel (see Fig. 7). The fabrication of the input channel modification is simple.
334
Such back-turning channels can be simply formed by cutting off some lower parts
335
of the adjacent walls of an input channel. The cold inflow will be provided with a
336
second pass through the channel and more heat exchanging area before entering
337
the catalyst bed. Thus the bed temperature difference will be determined by
338
the more uniform temperature gradient in the back-turning channel. Fig. 8
339
shows the different heat transfer behaviors and temperature profiles with and
Jo u
rn
al
Pr
331
340
with out back-turning channel (the figure is widened in y-direction to exhibit the
341
details of temperature profile). To demonstrate the effect of the back-turning
342
channel more intuitively, the axial distribution curves of channel temperature
343
are shown in Fig. 9. The maximal temperature difference in the single input
344
channel is about 220o C, while that in the back-turning input channel is about
345
120o C. Therefore by adding the back-turning channel, the channel temperature
346
difference can be nearly halved.
15
Page 25 of 36
Table 2: Geometric and operations parameters for monolithic reactors.
Vin (m/s)
y0,CH4
Conv.CH4
∆Tmax (o C)
Mei [45]
φ2.5mm × 2.54cm
427
1
0.68%
> 99%
150
Cominos [46]
φ1.2mm × 10cm
450
7.75
2%
81.5%
467
Hwang [47]
φ1.5mm × 8cm
594
7.3
3.5%
43%
396
f
Tin (o C)
3.4. Simulation results validation
oo
347
Channel size
This investigation has not been funded, so the experimental validations
349
may not be available at the present stage, but will be the priority in the future
350
works. Therefore, the simulation results can only be validated by the results of
351
some similar investigations.
pr
348
Although the flow pattern in the MPBBR is the axial free flow in the
353
channel and radial infiltration in the bed, the axial heat transfer behaviors is
354
similar to those in the monolithic reactors. The axial temperature distributions
355
in different monolithic reactors with coated Pt/alumina catalyst are shown in
356
Fig. 11. The corresponding geometric and operation parameters are listed in
357
table 2. The temperature distribution pattern in different monolithic reactors
358
are similar. A rapid growth of temperature exists near the entrance region and
359
the temperature gradient is small in the rest part. Although the combustion
360
temperature is low in the MPBBR due to the use of low temperature combustion
361
catalysts, the distribution pattern is similar. The difference is that the rapid
rn
al
Pr
e-
352
temperature growth is shifted into the input channel.
Jo u 362
363
Yan et al. [48] studied the heat transfer in a heat recirculation meso-
364
combustor (HRMC). The structure of the HRMC (Fig. 12) is similar to that of
365
the back-turning channel in the MPBBR. Yan et al. compared the temperature
366
profile between the HRMC and the conventional meso-combustor under the
367
conditions of Tin = 227o C, Vin = 0.5m/s, y0,CH4 = 3.5%, and the conversion
368
of methane is larger than 80%. Fig. 13 shows the temperature profiles at the
369
centerlines of the combustors. Due to the preheating inlet channels, the axial
370
temperature difference in the HRMC is only a quarter of that in the CMC.
16
Page 26 of 36
Table 3: Geometric parameters for different channel density.
cpsi
2h(mm)
Nb
ww (mm)
fw (v%)
fc (v%)
a
29
4
3
0.73
28.5
42.9
b
50
3
4
0.6
30.5
46.3
c
79
2.4
5
0.46
29.6
50.2
d
117
2
6
0.35
27.6
54.3
e
95
2
6
0.6
40.8
oo
f
Structure No.
44.4
Comparing Fig. 13 and Fig. 9, we can see similar heat transfer enhancement
372
characteristics.
373
3.5. Optimization of the geometric parameters
e-
pr
371
Now that we have proved the benefits of adding back-turning channels,
375
we can ulteriorly optimize the geometric parameters for the MPBBR elemen-
376
tary unit for 1v% & 0.5v% CH4 . For 0.5% concentration of methane, the
377
inlet temperature is set to be 405o C. The optimization should also consider
378
the manufacture feasibilities. The size of honeycomb monoliths are commonly
379
denoted by the channels per square inch (cpsi). Table 3 shows the geometric
380
parameters of some commercially available honeycomb monoliths. Channels
381
thinner than 2mm are not considered because the catalyst particle diameter is
382
0.18mm, whereas the channel width to diameter ratio should be at least 10:1.
383
The wall volume fraction fw and the catalyst packing fraction fc are calculated
Jo u
rn
al
Pr
374
384
by Eq. (20) and Eq. (21) respectively, based on the geometric relations denonted
385
in Fig. 14.
(2h + ww )2 − (2h)2 (2h + ww )2
(20)
hwb (h + ww /2)(2h + ww )(N + 2)
(21)
fw =
386
fc = 387
17
Page 27 of 36
388
Simulations are carried out to give the relations between the maximal tem-
389
perature difference and bed length for different channel density in Fig. 15 &
390
16. For the combustion of 1v% methane, it can be concluded from relation
392
curves a, b, c, d in Fig. 15 that with higher wall volume fraction, the maximal
393
temperature difference is reduced. However, the rule is not followed by structure
394
e. In addition, the maximal temperature differences seem to grow much faster
395
with L than expected for structure c and d. This is because thinner channel
396
width will increase the value of Pr and reduce the channel P eclet N umber.
397
Higher Pr will result in non-uniformity of infiltration velocity and thus of reac-
398
tion load, whereas lower P eclet N umber means relatively stronger radial heat
399
conduction in the input channel and larger axial difference of channel temper-
400
ature. Thus in the perspective of uniformizing the temperature distribution,
401
structure b tends to the best choice.
e-
pr
oo
f
391
As mentioned earlier, the activity of P d/Cex Zr1−x O2 catalyst will decrease
403
above 580o C and simulations show that when the maximal temperature differ-
404
ence in the bed exceed 50o C, the total methane conversion will be lower than
405
99%. Therefore, to be conservative, the threshold of the maximal temperature
406
difference is set to be 45o C.
al
Pr
402
From the dashed line of 45o C in Fig. 15, the maximal bed length for
408
different channel densities can be read. Although the catalyst packing fraction
409
is higher for structure c and d, the bed length cannot exceed 8.5cm, which
Jo u
rn
407
410
is not favorable to the handling capacity and manufacture feasibility (most
411
commercially available monoliths are longer than 10cm). Thus structure b is
412
finally chosen as the optimal channel density for the combustion of 1v% methane
413
and the maximal bed length can reach 12.2cm.
414
For the combustion of 0.5v% methane, from the relation curves in Fig. 16, it
415
can be seen that structure a, instead of b, has the best heat transfer performance.
416
This is because the heat effect is reduced due to lower concentration of methane
417
and the temperature distribution is mainly determined by the value of Pr . Thus,
418
for sturcture a, the maximal bed length can reach about 24cm. However, the 18
Page 28 of 36
catalyst packing fraction of structure a is the lowest, whereas our optimal design
420
principle is to achieve as high packing fraction as possible. Therefore, from the
421
dashed line in Fig. 16, structure c is finally chosen as the optimal channel
422
density for the combustion of 0.5v% methane and the maximal bed length can
423
reach 11.6cm.
424
3.6. Performance comparison between the MPBBR and CFRR
f
419
Detailed performance comparisons between CFRR and MPBBR for the
426
combustion of 1v% methane are listed in Table 4 and the detailed temperatuer
427
profiles in the MPBBR elementary unit are shown in Fig. 17. Due to the
428
distributed flow pattern in the MPBBR, the pressure drop doesn’t exceed that in
429
the CFRR, despite that the catalyst beds are packed with 0.18mm fine particles.
430
Since the catalyst effectiveness and catalyst packing fraction are both doubled,
431
the handling capacities per unit mass of catalyst and unit volume of reactor are
432
significantly increased. In this situation, to achieve the same handling capacity,
433
the MPBBR takes only half mass of catalysts and one third volume of reactor
434
used in the CFRR. For the heat transfer performance, the MPBBR realizes the
435
auto-thermal function without any external heat removal units and the maximal
436
temperature difference is maintained under 45 o C. More importantly, such high
437
performance catalytic combustion reactor of lean methane can be converted to
438
a compact modular mobile device to meet the demands of various scales of
439
exhaust gas treatment.
Jo u
rn
al
Pr
e-
pr
oo
425
440
4. Conclusion
441
The FBBR is a novel concept packed bed reactor with hierarchical channel
442
network and distributed flow pattern, which is designed for fast catalytic reac-
443
tions. The MPBBR is a representative implementation of the FBBR with the
444
capability of enhancing heat transfer performance for highly exothermic reac-
445
tions. The distribution uniformity of flow and temperature is the main challenge
446
of the optimal design. In this work, the heat transfer behaviors are qualitatively
19
Page 29 of 36
Table 4: Performance comparisons between MPBBR and CFRR [49].
MPBBR
Methane conversion (%)
>99
99.4
Catalyst particle diameter (mm)
7.5
0.18
Catalyst effectiveness (%)
42.1
90.1
Catalyst packing fraction (%)
>25
46.3
Maximal temperature difference in the bed (o C)
200
44.3
Total pressure drop (P a)
2500
2030
Unit mass handling capacity (kg CH4/kg Cat · h)
0.095
0.201
Unit volume handling capacity (kg CH4/m3 Cat · h)
14.7
50.7
pr
oo
f
CFRR
analyzed. In the input channel, the axial heat convection should be enhanced
448
while the radial heat conduction should be weakened to reduce the difference
449
in the bed entrance temperature. In the packed bed, the axial heat conduc-
450
tion should be enhanced to match the generation rate of reaction heat, whereas
451
the radial heat conduction should be dominant comparing to the radial heat
452
convection to realize the auto-thermal function.
Pr
e-
447
In the application case, numerical simulations give quantitative analysis
454
for the influence of channel structure and geometric parameters on the heat
455
transfer performance. The modified input channel with a back-turning chan-
456
nel can effectively uniformize the bed temperature distribution by reducing the
457
axial difference of channel temperature. The optimal channel density and ge-
458
ometric parameters are determined as 50 cpsi, and 2h = 3mm, N = 4, ww =
459
0.6mm, L = 12.2cm for the combustion of 1v% methane, and 79 cpsi, and
460
2h = 2.4mm, N = 5, ww = 0.46mm, L = 11.6cm for the combustion of 0.5v%
461
methane. The performance comparison between the CFRR demonstrates that
462
only half mass of catalysts and one third volume of reactor are needed based on
463
unit handling capacity. Therefore, with the flow and temperature distribution
464
uniformity under control, the MPBBR is proved to be an effective and economi-
465
cal alternative for many fast catalytic reactions. Further investigations and case
Jo u
rn
al
453
20
Page 30 of 36
466
studies will be carried out upon more industrial processes.
467
Acknowledgements
468
This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors.
470
References
oo
472
[1] W.-L. Liu, X. Li, Flow distribution in a novel fixed bed bionic reactor, Chemical Engineering Science 191 (2018) 525–536.
pr
471
f
469
[2] D. Vervloet, F. Kapteijn, J. Nijenhuis, J. R. van Ommen, Process intensifi-
474
cation of tubular reactors: Considerations on catalyst hold-up of structured
475
packings, Catalysis today 216 (2013) 111–116.
e-
473
[3] C. G. Visconti, G. Groppi, E. Tronconi, Highly conductive packed foams:
477
A new concept for the intensification of strongly endo-and exo-thermic
478
catalytic processes in compact tubular reactors, Catalysis Today 273 (2016)
479
178–186.
al
Pr
476
[4] C. Busse, H. Freund, W. Schwieger, Intensification of heat transfer in cat-
481
alytic reactors by additively manufactured periodic open cellular structures
482
(pocs), Chemical Engineering and Processing-Process Intensification 124
483
(2018) 199–214.
Jo u
rn
480
484
485
[5] J. Prins, J. Mulder, J. Schenk, Heat transfer in laminary flow between parallel plates, Applied Scientific Research 2 (1) (1951) 431.
486
[6] R. Cess, E. Shaffer, Heat transfer to laminar flow between parallel plates
487
with a prescribed wall heat flux, Applied Scientific Research, Section A
488
8 (1) (1959) 339–344.
489
[7] R. Terrill, Heat transfer in laminar flow between parallel porous plates,
490
International Journal of Heat and Mass Transfer 8 (12) (1965) 1491–1497.
21
Page 31 of 36
491
[8] V. Yeroshenko, L. Zaichik, B. Y. Bakhvalov, Heat transfer in laminar plane
492
channel flow with uniform suction or injection, International Journal of
493
Heat and Mass Transfer 24 (10) (1981) 1649–1655. [9] O. Cekmer, M. Mobedi, B. Ozerdem, I. Pop, Fully developed forced convec-
495
tion heat transfer in a porous channel with asymmetric heat flux boundary
496
conditions, Transport in porous media 90 (3) (2011) 791–806.
f
494
[10] P. Mondal, Thermodynamically consistent limiting forced convection heat
498
transfer in a asymmetrically heated porous channel: An analytical study,
499
Transport in porous media 100 (1) (2013) 17–37.
pr
oo
497
[11] E. V. Rebrov, J. C. Schouten, M. H. De Croon, Single-phase fluid flow
501
distribution and heat transfer in microstructured reactors, Chemical Engi-
502
neering Science 66 (7) (2011) 1374–1393.
e-
500
[12] J. Haber, M. Kashid, A. Renken, L. Kiwi-Minsker, Heat management in
504
single and multi-injection microstructured reactors: scaling effects, stability
505
analysis, and role of mixing, Industrial & Engineering Chemistry Research
506
51 (4) (2011) 1474–1489.
al
Pr
503
[13] E. N. Fuller, P. D. Schettler, J. C. Giddings, New method for prediction of
508
binary gas-phase diffusion coefficients, Industrial & Engineering Chemistry
509
58 (5) (1966) 18–27.
rn
507
[14] D. Gartling, C. Hickox, R. Givler, Simulation of coupled viscous and porous
511
flow problems, International Journal of Computational Fluid Dynamics
512
7 (1-2) (1996) 23–48.
Jo u
510
513
514
515
516
[15] W. J. Layton, F. Schieweck, I. Yotov, Coupling fluid flow with porous media flow, SIAM Journal on Numerical Analysis 40 (6) (2002) 2195–2218. [16] G. S. Beavers, D. D. Joseph, Boundary conditions at a naturally permeable wall, Journal of fluid mechanics 30 (1) (1967) 197–207.
22
Page 32 of 36
517
[17] B. Goyeau, D. Lhuillier, D. Gobin, M. Velarde, Momentum transport at
518
a fluid–porous interface, International Journal of Heat and Mass Transfer
519
46 (21) (2003) 4071–4081. [18] P. Schmitz, M. Prat, 3-d laminar stationary flow over a porous surface
521
with suction: Description at pore level, AIChE Journal 41 (10) (1995)
522
2212–2226.
f
520
524
[19] R. B. Bird, W. E. Stewart, E. N. Lightfoot, Transport phenomena john
oo
523
wiley & sons, New York 413.
[20] S. Jones, et al., Using the inertial coefficient, b, to characterize heterogene-
526
ity in reservoir rock, in: SPE Annual Technical Conference and Exhibition,
527
Society of Petroleum Engineers, 1987.
531
532
533
e-
Pr
530
ical engineering research and design 85 (9) (2007) 1245–1252. [22] G. R. Hadley, Thermal conductivity of packed metal powders, International Journal of Heat and Mass Transfer 29 (6) (1986) 909–920.
al
529
[21] J. Delgado, Longitudinal and transverse dispersion in porous media, Chem-
[23] P. Saffman, A theory of dispersion in a porous medium, Journal of Fluid Mechanics 6 (3) (1959) 321–349.
rn
528
pr
525
[24] S. M. Dommeti, V. Balakotaiah, D. H. West, Analytical criteria for validity
535
of pseudohomogeneous models of packed-bed catalytic reactors, Industrial
Jo u
534
536
& engineering chemistry research 38 (3) (1999) 767–777.
537
[25] D. Schlereth, O. Hinrichsen, A fixed-bed reactor modeling study on the
538
methanation of co2, Chemical Engineering Research and Design 92 (4)
539
(2014) 702–712.
540
[26] M. Halabi, M. De Croon, J. Van der Schaaf, P. Cobden, J. Schouten,
541
Modeling and analysis of autothermal reforming of methane to hydrogen
542
in a fixed bed reformer, Chemical Engineering Journal 137 (3) (2008) 568–
543
578. 23
Page 33 of 36
544
[27] A. Srinivasan, C. Depcik, One-dimensional pseudo-homogeneous packed-
545
bed reactor modeling: I. chemical species equation and effective diffusivity,
546
Chemical Engineering & Technology 36 (1) (2013) 22–32. [28] R. Ghidossi, J. Daurelle, D. Veyret, P. Moulin, Simplified cfd approach of
548
a hollow fiber ultrafiltration system, Chemical Engineering Journal 123 (3)
549
(2006) 117–125.
f
547
551
[29] A. Pak, T. Mohammadi, S. Hosseinalipour, V. Allahdini, Cfd modeling of
oo
550
porous membranes, Desalination 222 (1-3) (2008) 482–488.
[30] K. Damak, A. Ayadi, B. Zeghmati, P. Schmitz, A new navier-stokes and
553
darcy’s law combined model for fluid flow in crossflow filtration tubular
554
membranes, Desalination 161 (1) (2004) 67–77.
e-
pr
552
[31] G. Arzamendi, P. Di´eguez, M. Montes, J. Odriozola, E. F. Sousa-Aguiar,
556
L. Gand´ıa, Methane steam reforming in a microchannel reactor for gtl
557
intensification: A computational fluid dynamics simulation study, Chemical
558
Engineering Journal 154 (1-3) (2009) 168–173.
Pr
555
[32] S. Cruz, O. Sanz, R. Poyato, O. Laguna, F. J. Echave, L. Almeida, M. Cen-
560
teno, G. Arzamendi, L. Gandia, E. Souza-Aguiar, et al., Design and testing
561
of a microchannel reactor for the prox reaction, Chemical engineering jour-
562
nal 167 (2-3) (2011) 634–642.
rn
[33] K. S. Patel, A. K. Sunol, Modeling and simulation of methane steam re-
Jo u
563
al
559
564
forming in a thermally coupled membrane reactor, International Journal of
565
Hydrogen Energy 32 (13) (2007) 2344–2358.
566
[34] S. Kumar, S. Kumar, J. K. Prajapati, Hydrogen production by partial
567
oxidation of methane: modeling and simulation, international journal of
568
hydrogen energy 34 (16) (2009) 6655–6668.
569
[35] S. Salomons, R. Hayes, M. Poirier, H. Sapoundjiev, Flow reversal reactor
570
for the catalytic combustion of lean methane mixtures, Catalysis Today
571
83 (1-4) (2003) 59–69. 24
Page 34 of 36
[36] R. Litto, R. Hayes, H. Sapoundjiev, A. Fuxman, F. Forbes, B. Liu,
573
F. Bertrand, Optimization of a flow reversal reactor for the catalytic com-
574
bustion of lean methane mixtures, Catalysis today 117 (4) (2006) 536–542.
575
[37] S. Specchia, F. Conti, V. Specchia, Kinetic studies on pd/ce x zr1- x o2
576
catalyst for methane combustion, Industrial & Engineering Chemistry Re-
577
search 49 (21) (2010) 11101–11111.
f
572
[38] R. Abbasi, L. Wu, S. Wanke, R. Hayes, Kinetics of methane combustion
579
over pt and pt–pd catalysts, Chemical Engineering Research and Design
580
90 (11) (2012) 1930–1942.
pr
oo
578
[39] E. Garcıa, M. Osendi, P. Miranzo, Thermal diffusivity of porous cordierite
582
ceramic burners, Journal of applied physics 92 (5) (2002) 2346–2349.
583
[40] G. Neuer, R. Brandt, K. D. Maglic, N. Milosevic, G. Groboth, S. Rudtsch,
584
Thermal diffusivity of the candidate standard reference material cordierite,
585
High Temperatures–High Pressures(UK) 31 (5) (1999) 517–524.
Pr
e-
581
[41] X.-H. Han, Q. Wang, Y.-G. Park, C. Tjoen, A. Sommers, A. Jacobi, A
587
review of metal foam and metal matrix composites for heat exchangers and
588
heat sinks, Heat Transfer Engineering 33 (12) (2012) 991–1009.
rn
al
586
[42] C.-Y. Liu, W.-H. Tuan, S.-C. Chen, Preparation of porous sic ceramics for
590
thermal dissipation purposes, Ceramics International 41 (3) (2015) 4564–
591
4568.
Jo u
589
592
[43] G. Groppi, A. Belloli, E. Tronconi, P. Forzatti, A comparison of lumped and
593
distributed models of monolith catalytic combustors, Chemical Engineering
594
Science 50 (17) (1995) 2705–2715.
595
[44] N. C. Acikbas, Y. Ture, E. Gurlek, S. Ozcan, S. Soylu, G. Acikbas, T. Gudu,
596
Microstructural characterization, mechanical, physical and thermal prop-
597
erties of a diesel particulate filter, Arabian Journal for Science and Engi-
598
neering 43 (3) (2018) 1383–1394.
25
Page 35 of 36
599
[45] M. Hong, L. Chengyue, L. Hui, J. Shengfu, Simulation of catalytic com-
600
bustion of methane in a monolith honeycomb reactor, Chinese Journal of
601
Chemical Engineering 14 (1) (2006) 56–64. [46] V. Cominos, A. Gavriilidis, Theoretical investigation of axially non-uniform
603
catalytic monoliths for methane combustion, Chemical engineering science
604
56 (11) (2001) 3455–3468.
f
602
[47] C. Hwang, C. Lee, K. Lee, Numerical investigation on combustion charac-
606
teristics of methane in a hybrid catalytic combustor, Fuel 83 (7-8) (2004)
607
987–996.
pr
oo
605
[48] Y. Yan, S. Feng, Z. Huang, L. Zhang, W. Pan, L. Li, Z. Yang, Thermal
609
management and catalytic combustion stability characteristics of premixed
610
methane/air in heat recirculation meso-combustors, International Journal
611
of Energy Research 42 (3) (2018) 999–1012.
Pr
e-
608
[49] S. Salomons, R. Hayes, M. Poirier, H. Sapoundjiev, Modelling a reverse
613
flow reactor for the catalytic combustion of fugitive methane emissions,
614
Computers & chemical engineering 28 (9) (2004) 1599–1610.
Jo u
rn
al
612
26
Page 36 of 36