You are on page 1of 9

Fuel Processing Technology 125 (2014) 139147

Contents lists available at ScienceDirect

Fuel Processing Technology


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

Kinetic modeling of methanol synthesis over commercial catalysts based


on three-site adsorption
Nonam Park a,b, Myung-June Park a,b,, Yun-Jo Lee c, Kyoung-Su Ha c,1, Ki-Won Jun c,
a
b
c

Department of Chemical Engineering, Ajou University, Suwon 443-749, Republic of Korea


Department of Energy Systems Research, Ajou University, Suwon 443-749, Republic of Korea
Research Center for Green Catalysis, Korea Research Institute of Chemical Technology (KRICT), Daejeon 305-600, Republic of Korea

a r t i c l e

i n f o

Article history:
Received 6 December 2013
Received in revised form 26 March 2014
Accepted 31 March 2014
Available online xxxx
Keywords:
Methanol synthesis
Kinetic model
Three-site adsorption
Parameter estimation
Effectiveness factor

a b s t r a c t
A kinetic model for the synthesis of methanol over commercial catalysts was developed based on the adsorption
of CO and CO2 onto various active sites of Cu, while the kinetic parameters were estimated by tting 118 experimental sets of data under a variety of conditions. When both CO and CO2 hydrogenations take place, CO conversion was inuenced by the change in temperature as well as in space velocity and pressure, while CO2 conversion
was minimally inuenced as a result of its correlation to the watergas shift reaction. However, the CO and H2
fractions signicantly inuenced both conversions. The analysis based on the model developed in this investigation clearly conrmed that the contribution of each reaction and the maximum methanol concentration was
expressed as a function of temperature and the CO fraction. Additionally, catalysts with varying particle sizes
were used to evaluate the effect of the internal diffusion limitation on catalytic performance and the effectiveness
factors were subsequently calculated and regressed with respect to the particle size.
2014 Elsevier B.V. All rights reserved.

1. Introduction
Methanol (MeOH) is a key industrial chemical that can be used either as a solvent or a fuel, in addition to being able to be readily converted into other useful products such as formaldehyde, amines, acetic acid,
and esters [1]. MeOH is also used to produce ethylene or propylene in
the methanol-to-olens process and is capable of conveniently storing
and transporting CO and H2 [2,3]. While fossil fuels have contributed
substantially to atmospheric pollution by emitting signicant amounts
of greenhouse gases, particularly CO2 [4], recycling excess CO2 from industrial waste and the atmosphere and reacting it with H2 to produce
MeOH could potentially mitigate the global warming problem. Additionally, MeOH is directly oxidized in air (without H2) to CO2 and H2O
to produce electricity in direct-MeOH fuel cells [5].
Since the reaction employed in the synthesis of MeOH is highly
exothermic, the heat released from the reaction needs to be efciently
removed to keep the reactor isothermal. For this reason, much research
is devoted to kinetic and reactor modeling when designing commercial
processes for mass production [3]. Previously reported kinetic models
include mechanistic kinetic models based on detailed reaction

Correspondence to: M.-J. Park, Department of Chemical Engineering, Ajou University,


Suwon 443-749, Republic of Korea. Tel.: +82 31 219 2383; fax: +82 31 219 1612.
Corresponding author. Tel.: +82 42 860 7671; fax: +82 42 860 7388.
E-mail addresses: mjpark@ajou.ac.kr (M.-J. Park), kwjun@krict.re.kr (K.-W. Jun).
1
Present address: Department of Chemical and Biomolecular Engineering, Sogang
University, Seoul 121-742, Republic of Korea.

http://dx.doi.org/10.1016/j.fuproc.2014.03.041
0378-3820/ 2014 Elsevier B.V. All rights reserved.

mechanisms [69] and the power law model [10,11]. Grabow and
Mavrikakis [12] developed a microkinetic model for the synthesis of
MeOH, including the watergas shift (WGS) reaction using density
functional theory calculations, while Ovesen et al. [13] described a static
and dynamic (changes in particle shape and active surface area)
microkinetic model for the synthesis of MeOH over a Cu/ZnO catalyst.
Both CO and CO2 hydrogenations can be used in the synthesis of
MeOH to achieve high productivity, and Klier et al. [14] proposed
three models based on different assumptions concerning how CO, CO2,
and H2 compete for the active sites. McNeil et al. [15] assumed that
the adsorption of hydrogen occurs on ZnO in contrast to the adsorption
of CO and CO2 on Cu1+ and Cu0, respectively, and it was also determined
that the catalyst has both reduced and oxidized (such as Cu2+ which
may be formed by the oxidation of Cu metal) sites for CO and CO2 hydrogenations, respectively, in the presence of an optimal amount of
CO2 or H2O [16]. When CO and CO2 hydrogenations are occurring simultaneously, the MeOH production rate reaches a maximum and subsequently decreases as the CO2 concentration increases, indicating that
CO2 inhibits CO hydrogenation [17], while the synergetic effects of CO2
had been suggested in our previous work [8].
However, despite many useful kinetic models for MeOH synthesis
(as discussed above, and as shown in the list of models in the literature
[18,19]), very few investigations have addressed the development of kinetic models based on three-site adsorption (Cu1 +, Cu0, and ZnO).
Therefore, detailed elementary steps for both CO and CO2 hydrogenations based on two-site adsorption [19] were modied by considering
additional adsorption site for CO2 in our previous work [8], and the

140

N. Park et al. / Fuel Processing Technology 125 (2014) 139147

model for the three-site adsorption was developed for the purpose of its
application to the catalysts developed in the Korea Research Institute of
Chemical Technology (KRICT). It is worth noting that the KRICT catalysts
exhibited different rate determining step (RDS) (the surface reaction of
a methoxy species, the hydrogenation of the formate intermediate
HCO2, and the formation of a formate intermediate for CO and CO2 hydrogenations and WGS reaction), relative to the reported model for
commercial catalysts [19]. Therefore, since commercial catalysts were
used in the present study because of their use in large scale commercial
reactors, the elementary steps in our previous work [8] were considered
while the RDS assumption for commercial catalyst [19] was applied. In
such a way, the model for the three-site adsorption was developed in
the present study.

detected with the thermal conductivity detector and a HP-PLOT Q capillary column coupled to ame ionization detector was used to separate
and analyze all hydrocarbons, including MeOH.
3. Results and discussion
3.1. Reaction rates
The synthesis of MeOH was comprised of three main reactions: (1)
CO hydrogenation, (2) the WGS reaction, and (3) CO2 hydrogenation.
Additionally, the dehydration of MeOH to produce dimethyl ether
(DME) was considered in this study.
CO 2H2 CH3 OH

CO2 H2 CO H2 O

CO2 3H2 CH3 OH H2 O

2CH3 OHCH3 OCH3 H2 O:

2. Experimental
A commercial catalyst (Cu/ZnO/Al2O3, Sd-Chemie, MegaMax 700)
is used for the kinetic experiment regarding the synthesis of MeOH.
The pelletized catalyst (5 mm diameter; 3 mm height) was broken
and sieved to produce three samples of uniform size: 0.15 mm
0.25 mm, 0.75 mm0.85 mm, and 1.5 mm2.5 mm; which were denoted as S1, S2, and S3, respectively. Additionally, the original catalyst pellet, denoted as S4, was also used without breaking. The S1 catalyst was
used for the kinetic investigation, while the others were used to evaluate the effects of the particle sizes on the catalytic performance.
The kinetic data was collected using a continuous tubular ow xedbed microreactor (see Fig. 1 for the schematic of the experimental system). The temperature within the reactor was controlled by adjusting
the furnace temperature and the ow rate was controlled by a mass
ow controller. The pressure was precisely controlled by a back pressure regulator and monitored by a digital pressure sensor. The catalyst
samples (0.4 g) were diluted with similar-sized and inert -Al2O3 particles (1.2 g) and packed together into a stainless steel reactor (internal
diameter = 7 mm). The catalyst was reduced under a ow of H2
(100 mL/min) and temperature was increased from room temperature
to a predetermined point at a rate of 2 K/min and held for 3 h. After
reduction, the pressure was increased according to the operating conditions, and the H2 gas was then replaced by the gas used in the synthesis
of MeOH. Detailed conditions regarding temperature, pressure, and feed
gas composition are listed in Table 1. The reaction products were analyzed by an on-line gas chromatograph (Young Lin), where a Carboxen
1000 column was installed to separate the CO, CO2, H2, and Ar that were

All reactions occurred simultaneously on the surface of the catalyst


and CO and CO2 were assumed to adsorb on the various sites of catalyst
during the synthesis of MeOH [8,14]. Table 2 displays the elementary reactions for CO and CO2 hydrogenations and the WGS reaction, where
the adsorption site of CO2 in the present study was denoted as s3
(s1 was used to denote both CO and CO2 hydrogenations in the
literature [19]). It is worth noting that the dehydration of methanol to
dimethyl ether was included since experimental data showed the production of DME, while other byproducts such as higher alcohol, methane and hydrocarbons were detected with trace (negligible) amount.
Therefore, although it is well-known that methane and other hydrocarbons could appear as byproducts under certain reaction conditions and
catalysts, those reactions were not considered in the present study.
The elementary steps of (A3), (B2), and (C3) were chosen to be the
RDSs for CO hydrogenation, WGS reaction, and CO2 hydrogenation,
respectively, as determined in the literature [19], and all the other
steps were assumed to be at equilibrium. The adsorption of MeOH
was assumed to be negligible, while the fractions of surface CO

Pressure gauge

CO/H2 / Ar
CO2 /H 2

Reactor

H 2 in He
Mass flow Check valve
controller
(MFC)

Safety valve
Filter

GC
PC System

Back pressure
regulator (BPR)

Fig. 1. Schematic of the experimental system used in this study.

Trap

N. Park et al. / Fuel Processing Technology 125 (2014) 139147

(CO(s1)), CO2 (CO2(s3)), hydrogen atoms (H(s2)), and water (H2O(s2))


were determined from the adsorption steps as follows:
COs1 K CO f CO s1

CO2s3 K CO2 f CO2 s3

from the literature [21]. Since only two of the four reaction rates are
independent from a thermodynamic standpoint, the KP,C can be
established as linear combinations of the other values.
h
i 1:183  104
2

ln K P;A bar
29:061
T K

Hs2 K H2 f H2 s2

ln K P;B 

H2Os2 K H2O f H2O s2 :

K P;C K P;A K P;B :

0:5 0:5

In the step (A3) of Table 2, the fractions of H2CO(s1) and H3CO(s1)


were calculated by inserting the above equations as follows:
2

H2COs1 K A1 K A2 COs1 Hs2 =s2 K CO K H2 K A1 K A2 s1 f CO f H2




0:5 0:5
H3COs1 s1 f CH3OH = K A4 K H2 f H2 :

r DME

Eqs. (9) and (10) were inserted into the rate of the (A3) step to obtain:
r CO



kA3 H2COs1 Hs2 H3COs1 s2 =K A3
!
f CH3OH
0
1:5
s1 s2 :
kA K CO f CO f H2
K P;A f 0:5
H2

16

4:773  103
4:672
T K

17

18

For the DME production, a reaction rate equation was available in


the literature [22] and used without alteration:

10

141

h
i
kDME K 2CH3OH C 2CH3OH C H2O C DME =K P;DME

4
p
1 2 K CH3OH C CH3OH K H2O;DME C H2O

19

where Ci was the concentration of component i in mol/m3, and the


reaction equilibrium constant was:
K P;DME 0:106 exp

!
2:1858  104
:
RT

20

11
3.2. Reactor model and parameter estimation

Site balance equations for site 1, 2 and 3 were calculated by assuming that the total number of sites per weight of catalyst was constant in
addition to neglecting terms originating from intermediate products
[19], as follows:
1
1
; s2
;
0:5
1 K CO f CO
f
1 K 0:5
H2 H2 K H2O f H2O
1

:
1 K CO2 f CO2

s1

s3
12

By inserting Eqs. (12) to (11), the following CO hydrogenation reaction rate was obtained:

r CO

h

i
0:5
k0A K CO f CO f 1:5
H2 f CH3OH = K P;A f H2

:

0:5
1 K CO f CO 1 K 0:5
H2 f H2 K H2O f H2O

13

The dimensionless Mears parameter [23,24] was calculated using


experimental data and physical properties to evaluate the contribution
of external mass diffusion. The values ranged from 2.34 102 to
3.17 102, which were below the threshold value of 0.15, indicating
negligible external mass diffusion. The occurrence of internal pore diffusion limitation was determined based on the WeiszPrater criterion [23,
25,26], where the existence of the concentration gradient was determined if the value of the dimensionless WeiszPrater parameter (CWP)
was much greater than one (CWP 1). Since the values of CWP in this
study were close to one during each experiment, no internal diffusion
limitations were detected. Therefore, a plug ow model without dispersion was used to simulate the reactor as follows [8]:
Mass balance :

us

NR
X
dC i
B
r i; j 0
dz
j1

21

In the same manner, the CO2 hydrogenation and the WGS reaction
rates were determined as follows:


COs3 H2Os2
r WGS kB2 HCO2s3 Hs2
K B2



 K f K
0:5
0:5
0:5 0:5
H2O f H2O s2
kB2 K CO2 K H2 K B1 f CO2 f H2 s3 K H2 f H2 s2 CO CO s3
K B2
h
i
k0B K CO2 f CO2 f H2 f CO f H2O =K P;B


14

0:5
1 K CO2 f CO2 1 K 0:5
H2 f H2 K H2O f H2O


H3CO2s3 s2
r CO2 kC3 H2CO2s3 Hs2
"
K C3
!
#


K
f H2O f CH3OH s3 s2
0:5 0:5
kC3 K CO2 K H2 K C1 K C2 f CO2 f H2 s3 K H2 f H2 s2 H2O
1:5
1:5
K H2 K C4 K C5 K C6 f H2 K C3
h

i
0
1:5
1:5
kC K CO2 f CO2 f H2 f H2O f CH3OH = K P;C f H2



15
0:5
1 K CO2 f CO2 1 K 0:5
H2 f H2 K H2O f H2O

where f denotes the fugacity, which was calculated using the generalized correlations for the fugacity coefcient in the literature [20], and
KP,i denotes the reaction equilibrium constant, which was obtained

Energy balance :

g us C P

Boundary conditions :

NR
X
dT
4U
H j r j
T T
B
dz
Dt w
j1

at

C i C i;in ;

z 0;

T T in :

22

23

The estimation of the parameters was based on minimizing the objective function (Fobj) and the residuals of square errors of the objective
elements (CO and CO2 conversions) were expressed as follows:
F obj

"
NE X
X
n

wi

X i;calc X i; exp
X i; exp

!2 #
24
n

where NE and wi represent the number of experimental conditions


and the weighting factor, respectively. It is important to note that
prior to the estimation, the experimental values of the CO and CO2 conversions were compared to the equilibrium conversions under the

142

N. Park et al. / Fuel Processing Technology 125 (2014) 139147

Table 1
Experimental conditions.
Case

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74

Pressure [bar]

50
50
50
50
50
50
50
50
50
50
70
90
50
50
50
50
50
50
50
50
50
50
50
50
50
50
50
70
70
70
70
70
70
70
70
70
70
70
70
70
70
50
50
50
50
60
60
60
60
70
70
70
70
80
80
80
80
50
50
50
50
50
50
50
50
50
50
50
50
50
50
50
50
50

Temp. [K]

523
543
573
593
613
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
543
543
543
543
543
543
523
523
523
523
523
523
523
523
523
523
523
523
523
533
543
543
503
513
523
533
543
553
563
573
583
593
613
503
513
523
533
543
553

Space velocity
[mL/(gcath)]

Feed composition
CO

CO2

H2

Ar

CO/(CO + CO2)

20,000
20,000
20,000
20,000
20,000
8000
20,000
30,000
40,000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
10,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000

0.19
0.19
0.19
0.19
0.19
0.19
0.19
0.19
0.19
0.19
0.19
0.19
0.00
0.11
0.17
0.22
0.26
0.27
0.28
0.29
0.32
0.18
0.16
0.13
0.09
0.19
0.14
0.14
0.19
0.00
0.17
0.22
0.27
0.29
0.32
0.00
0.17
0.22
0.27
0.29
0.32
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.07
0.07
0.07
0.07
0.19
0.19
0.19
0.19
0.19
0.19
0.19
0.19
0.19
0.19
0.19
0.11
0.11
0.11
0.11
0.11
0.11

0.11
0.11
0.11
0.11
0.11
0.11
0.11
0.11
0.11
0.11
0.11
0.11
0.24
0.16
0.11
0.07
0.05
0.03
0.02
0.02
0.00
0.10
0.09
0.07
0.05
0.11
0.08
0.08
0.11
0.24
0.11
0.07
0.03
0.02
0.00
0.24
0.11
0.07
0.03
0.02
0.00
0.24
0.22
0.17
0.12
0.24
0.22
0.17
0.12
0.24
0.22
0.17
0.12
0.05
0.05
0.05
0.05
0.11
0.11
0.11
0.11
0.11
0.11
0.11
0.11
0.11
0.11
0.11
0.06
0.06
0.06
0.06
0.06
0.06

0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.72
0.69
0.67
0.66
0.65
0.64
0.64
0.64
0.63
0.67
0.70
0.77
0.83
0.70
0.50
0.50
0.70
0.72
0.67
0.66
0.64
0.64
0.63
0.72
0.67
0.66
0.64
0.64
0.63
0.72
0.75
0.80
0.86
0.72
0.75
0.80
0.86
0.72
0.75
0.80
0.86
0.62
0.62
0.62
0.62
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.83
0.83
0.83
0.83
0.83
0.83

0
0
0
0
0
0
0
0
0
0
0
0
0.04
0.05
0.05
0.05
0.05
0.05
0.05
0.05
0.05
0.05
0.05
0.04
0.03
0
0.29
0.29
0
0.04
0.05
0.05
0.05
0.05
0.05
0.04
0.05
0.05
0.05
0.05
0.05
0.04
0.04
0.03
0.02
0.04
0.04
0.03
0.02
0.04
0.04
0.03
0.02
0.26
0.26
0.26
0.26
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.00
0.40
0.60
0.75
0.85
0.90
0.93
0.95
1.00
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.00
0.60
0.75
0.90
0.95
1.00
0.00
0.60
0.75
0.90
0.95
1.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.58
0.58
0.58
0.58
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64

H2/(2CO + 3CO2)

1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.17
1.65
2.50
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.16
1.60
2.39
1.00
1.16
1.60
2.39
1.00
1.16
1.60
2.39
2.04
2.04
2.04
2.04
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
2.00
2.00
2.00
2.00
2.00
2.00

Conversion[%]
CO

CO2

41.05
33.78
18.06
9.14
3.27
54.28
41.21
34.13
28.93
52.29
66.78
74.78
9.36
52.54
52.6
49.9
45.62
44.6
42.21
36.5
20
57.55
57.96
62.31
69.77
57.16
40.9
57.4
73.03
7.4
64.46
60.85
57.41
54.19
31.58
10.22
49.93
48.44
46.38
45.66
41.31
8.44
9.13
9.88
10.51
7.89
8.35
8.96
8.58
7.3
7.8
8.33
8.02
75.93
67.36
59.05
57.91
37.12
43.79
42.66
38.42
32.21
27.94
23.75
17.56
11.61
8.42
6.31
55.43
52.82
55.8
51.8
45.95
39.14

7.84
6.05
6.18
6.8
8.38
9.46
6.1
4.82
4.62
10.68
10.05
10.98
23.32
10.73
8.09
0.37
0.05
0
0
0
0.66
7.82
8
9.38
15.11
5.8
4.32
5.32
11.01
28.7
10.02
4.21
6.43
23.02
1.24
26.46
4.78
1.27
35.65
47.62
4.34
24.61
25.39
31.47
40.94
26.21
28
34.23
45.15
28.36
30.37
35.95
48.44
16.75
18.94
20.96
17.01
8.29
8.34
8.32
7.86
7.14
6.85
6.94
5.88
5.77
4.63
5.75
15.62
30.12
15.29
14.04
12.12
9.29

Close to the equil.

Remarks

Temperature

Space velocity

Pressure

CO/(CO + CO2)

H2/(2CO + 3CO2)

Inert gas

CO/(CO + CO2)

CO/(CO + CO2)

CO/(CO + CO2) = 0

CO/(CO + CO2) = 0

CO/(CO + CO2) = 0

Close to ICI plant


operating conditions

Temperature

Temperature

N. Park et al. / Fuel Processing Technology 125 (2014) 139147

143

Table 1 (continued)
Case

75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118

Pressure [bar]

50
50
50
50
50
50
50
50
50
50
50
50
50
50
70
70
70
70
70
70
70
70
70
70
70
70
50
70
90
50
50
50
50
50
50
50
50
50
50
50
50
50
50
50

Temp. [K]

563
573
503
513
523
543
503
513
523
533
543
553
563
573
523
493
503
513
523
533
543
553
573
523
573
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523
523

Space velocity
[mL/(gcath)]

Feed composition
CO

CO2

H2

Ar

CO/(CO + CO2)

20,000
20,000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
8000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
20,000
8000
8000
8000
20,000
20,000
20,000
20,000
20,000

0.11
0.11
0.19
0.19
0.19
0.19
0.11
0.11
0.11
0.11
0.11
0.11
0.11
0.11
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.18
0.00
0.11
0.17
0.22
0.27
0.29
0.32
0.18
0.13
0.09
0.18
0.16
0.13
0.11
0.09

0.06
0.06
0.11
0.11
0.11
0.11
0.06
0.06
0.06
0.06
0.06
0.06
0.06
0.06
0.10
0.10
0.10
0.10
0.10
0.10
0.10
0.10
0.10
0.10
0.10
0.10
0.10
0.10
0.10
0.24
0.16
0.11
0.07
0.03
0.02
0.00
0.10
0.07
0.05
0.10
0.09
0.07
0.06
0.05

0.83
0.83
0.70
0.70
0.70
0.70
0.83
0.83
0.83
0.83
0.83
0.83
0.83
0.83
0.67
0.67
0.67
0.67
0.67
0.67
0.67
0.67
0.67
0.67
0.67
0.67
0.67
0.67
0.67
0.72
0.69
0.67
0.66
0.64
0.64
0.63
0.67
0.77
0.83
0.67
0.70
0.77
0.80
0.83

0
0
0
0
0
0
0
0
0
0
0
0
0
0
0.05
0.05
0.05
0.05
0.05
0.05
0.05
0.05
0.05
0.05
0.05
0.05
0.05
0.05
0.05
0.04
0.05
0.05
0.05
0.05
0.05
0.06
0.05
0.04
0.03
0.05
0.05
0.04
0.03
0.03

0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.00
0.40
0.60
0.75
0.90
0.95
1.00
0.64
0.64
0.64
0.64
0.64
0.64
0.64
0.64

corresponding experimental conditions, and the values that deviated


from the equilibrium conversion by more than 15% were subsequently
used for the kinetic data. Thus, only 74 conditions (out of 118 total

H2/(2CO + 3CO2)

2.00
2.00
1.00
1.00
1.00
1.00
2.00
2.00
2.00
2.00
2.00
2.00
2.00
2.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.00
1.65
2.50
1.00
1.17
1.65
2.00
2.50

Conversion[%]
CO

CO2

32.7
26.75
56.42
55.65
51.37
36.57
76.18
72.28
65.93
59.27
50.32
44.14
39.76
36.69
54.23
31.72
55.62
57.92
53.21
49.76
44.3
37.45
27.95
53.18
26.22
53.73
41.26
48.59
53.21
15.13
33.55
35.57
38.76
30.45
28.93
13.76
53.92
64.84
71.93
30.36
31.59
36.83
38.71
40.1

7.42
3.88
12.15
11.28
10.09
9.22
18.88
17.86
17.07
15.9
11.59
10.24
6.94
2.51
5.45
4.09
5.61
5.61
3.98
3.72
3.35
1.58
0.31
4.23
2.89
5.18
6.37
4.28
3.98
25.64
8.64
4.23
6.29
11.95
15.01
0.82
7.47
17.21
24.26
7.16
8.85
14.02
17.64
20.91

Close to the equil.

Remarks

Temperature

Temperature

Temperature

Pressure

CO/(CO + CO2)

H2/(2CO + 3CO2)

H2/(2CO + 3CO2)

conditions) were selected for the estimation (the conditions marked


with solid circles in Table 1 were determined to be close to the equilibrium conversion and were excluded from the estimation). Additionally,

Table 2
Elementary reactions for synthesis of MeOH.

Adsorption

Elementary step

Equilibrium constant

CO + s1 CO(s1)

COs1
f CO s1

K CO2 f CO2s3s3
CO2

CO2 + s3 CO2(s3)
H2 + 2 s2 2H(s2)

Surface reaction
(A) CO hydrogenation reaction

(B) Watergas shift reaction


(C) CO2 hydrogenation reaction

K CO

K H2

2Hs2
f H2 2s2

H2O + s2 H2O(s2)

K H2O

Elementary steps
(A1): CO(s1) + H(s2) HCO(s1) + s2
(A2): HCO(s1) + H(s2) H2CO(s1) + s2
(A3): H2CO(s1) + H(S2) H3CO(s1) + s2
(A4): H3CO(S1) + H(s2) CH3OH + s1 + s
(B1): CO2(S3) + H(s2) HCO2(s3) + s2
(B2): HCO2(S3) + H(s2) CO(s3) + H2O(s2)
(C1): CO2(s3) + H(s2) HCO2(s3) + s2
(C2): HCO2(s3) + H(s2) H2CO2(s3) + s2
(C3): H2CO2(s3) + H(s2) H3CO2(s3) + s2
(C4): H3CO2(s3) + H(s2) H2CO(s3) + H2O(s2)
(C5): H2CO(s3) + H(s2) H3CO(s3) + s2
(C6): H3CO(s3) + H(s2) CH3OH + s3 + s2

Equilibrium constant
KA1
KA2
KA3
KA4
KB1
KB2
KC1
KC2
KC3
KC4
KC5
KC6

H2Os2
f H2O s2

144

N. Park et al. / Fuel Processing Technology 125 (2014) 139147

Table 3
Kinetic parameters estimated in the present study and reported valuesa.
Parameters

Units

Estimated
1.5

kA

mol/(kgcatsbar )

kB

mol/(kgcatsbar)

kC

mol/(kgcatsbar1.5)

kDME

mol/(kgcats)

KH2O

bar1

1:88  108 exp

Reported


1:40  10 exp

113;711
RT



1:16  10 exp 126;573
RT


7:08  104 exp 68;252
RT


8:54  109 exp 123;779
RT


3:80  1010 exp 80;876
RT

3:81  109 exp

10

2:27  103 exp


10

1:03  10





exp

113;045
RT
126;545
RT

68;345
RT

105;000
RT

References
[19]
[19]
[19]
[22]

R = 8.314 J/(molK).

the sensitivity analysis conrmed that the adsorption equilibrium constants do not signicantly impact the overall reaction rates; therefore,
the reported values [18,19,22] were used without modication, but
the adsorption coefcient of H2O (KH2O) was estimated because it was
not reported in the literature.
The estimation was conducted using the lsqcurvet subroutine in
MATLAB (MathWorks, Inc.), where the LevenbergMarquardt method
was applied and the estimated parameters are listed in Table 3. Fig. 2
displays the comparison between the experimental data and the simulated results with the corresponding statistics included and the effectiveness of the developed kinetic model was also shown. The mean of
absolute relative residuals (MARR) value for the CO2 conversion was
higher than that of CO conversion. This result was attributed to the
fact that the experimental values of the CO2 conversion under some
conditions were close to zero and the relative errors were thus calculated to be very high due to very small denominator values. Additionally,
when water and methanol in the reactor efuent were trapped to
prevent it from entering the GC column, a small amount of CO2 might
have remained dissolved in the liquid phase, resulting in increased measurement errors [27]. It was hard to directly measure dissolved CO2 in a
blank test because a minimal amount of CO2 was dissolved in the liquid
phase compared to that in the gas phase (but large enough to affect the
measurement errors of the CO2 conversions). We instead calculated the
weight fractions of CO2 in the liquid phase using a process simulator to

(a)

(1) (2) (3)

CO conversion [%]

100

(4)

(5) (6) (7)

determine the amount of CO2 that can be dissolved in the liquid phase as
a result of condensation. The values ranged from 0.14 wt.% to 6.74 wt.%
under all operating conditions, indicating that the dissolved CO2 varies
signicantly with respect to the experimental conditions which results
in high individual errors in some cases.
The activation energies of the estimated kinetic parameters were
nearly the same as the values reported in the literature [19], with the exception of the DME production rate constant being approximately 17%
higher than the value reported in the literature [22]. Meanwhile, the
pre-exponential factor of kC was determined to be approximately 30
times higher than the previously reported value, indicating that the
CO2 hydrogenation rates were relatively high. This result was conrmed
by the conditions in groups (9)(11) of Fig. 2 (experimental conditions
4253 in Table 1), and the CO2 conversion was determined to be
sufciently high. It is also important to note that the CO conversion
exhibited slightly negative values as a result of the weak reverse WGS
reaction rate and the CO2 was almost entirely consumed by the CO2
hydrogenation.
Although much effort was devoted towards maintaining the isothermal condition within the reactor in this study, there was still the possibility of a temperature gradient existing due to the high exothermicity
of the reaction; therefore, the energy balance equation (Eq. (22)) was
taken into consideration. It has been previously reported that the overall
heat transfer coefcient in the catalytic bed typically falls in the range of

(8) (9) (10)(11)(12)

(13)

(14) (15) (16)

(17)

(18) (19) (20) (21)

80
60
40
20
0
-20
0

10

20

30

40

50

60

70

80

90

100

110

120

CO2 conversion [%]

(b)
60
experiment
experiment
simulation
simulation

40
20
0
-20
-40
0

10

20

30

40

50

60

70

80

90

100

110

120

Case
Fig. 2. Comparison of the (a) CO and (b) CO2 conversions between the experimental and simulated data. The mean of the absolute relative residuals was 9.40% and 72.12% for the CO and
CO2 conversions, respectively. Numbers in parentheses above the diagram correspond to the effects of various operating conditions (detailed conditions are listed in Table 1): effects of
(1) temperature; (2) space velocity; (3) pressure; (4) CO fraction at 50 bar and 250 C; (5) H2 fraction; (6) inert gas fraction; CO fraction at 70 bar, (7) 250 C and (8) 270 C; H2/CO2
ratio with no CO for (9) 50 bar, (10) 60 bar and (11) 70 bar; (12) temperature and space velocity (close to ICI plant operating conditions); temperature with H2/(2CO + 3CO2) and SV
[mL/(gcath)] (13) 1 and 20000; (14) 2 and 20000; (15) 1 and 8000; and (16) 2 and 8000; (17) temperature at 70 bar; (18) pressure; (19) CO fraction; H2 fraction with SV of (20)
8000 and (21) 20,000.

(a)

(b)

MeOH concentration [mol/m3]

Accumulated rCO [mol/kgcat/s]

N. Park et al. / Fuel Processing Technology 125 (2014) 139147

160
140
120
100
80
60
40
20
0
540

Te

mp 530
era 520
tur
e [ 510 500
K]

0.8

1.0

0.6

CO f

0.4

o
racti

0.2

3.0e-4
2.5e-4
2.0e-4
1.5e-4
1.0e-4
5.0e-5
0.0

540

Te 530
mp
era 520
tur 510
e[
K] 500 1.0

0.8

0.6

0.4

0.2

on
racti
CO f

(d)
2.5e-4
2.0e-4
1.5e-4
1.0e-4
5.0e-5
0.0
-5.0e-5

540

Te
mp 530
era 520
tur
e [ 510
K]
500
1.0

0.8

0.6

0.4

0.2

on
racti
CO f

Accumulated rCO2 [mol/kgcat/s]

Accumulated rWGS [mol/kgcat/s]

3.5e-4

(c)

145

3.0e-4
2.5e-4
2.0e-4
1.5e-4
1.0e-4
5.0e-5
0.0

540

Te 530
mp
era 520
tur 510
e[
K] 500 1.0

0.8

0.6

0.4

0.2

on
racti
CO f

Fig. 3. Effects of temperature and CO fraction (=CO/(CO + CO2)) on (a) the MeOH concentration upon exiting the reactor, and the accumulated rates of (b) the CO hydrogenation, (c) the
WGS reaction, and (d) the CO2 hydrogenation. The pressure, space velocity, and H2 fraction (=H2/(2CO + 3CO2)) were specied as 50 bar, 8000 mL/(gcat h) and 1.0, respectively.

100300 W/(m2K) [2830], and the heat transfer coefcient that was
also estimated was determined to be 143 W/(m2K).
It should be noticed that, although the model developed in the
presented study is based on the intrinsic kinetics, its application is
limited to the catalyst with similar composition. This is attributable to
the fact that different composition may have different RDS, as discussed
in the introduction section (RDS's of KRICT and commercial catalysts
were different).

(a)
Effectiveness factor

1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.0

exp (523 K)
exp (533 K)
regression (523 K)
regression (533 K)

0.5

1.0

1.5

2.0

2.5

3.0

(b)
CO conversion [%]

60
exp (523 K)
exp (533 K)
sim (523 K)
sim (533 K)

50
40
30
20
10
0.0

0.5

1.0

1.5

2.0

2.5

3.0

Particle radius [mm]


Fig. 4. (a) Effectiveness factor calculated using experimental data and regressed lines:
y = 1.019 exp( 262.2x) for 523 K (R2 = 0.9552) and y = 1.034 exp( 250.4x) for
533 K (R2 = 0.9692); and (b) the CO conversion for various particle sizes and temperatures
(Catalyst weight 0.4 g, inert weight = 1.2 g, P = 50 bar, SV = 20,000 mL/(gcat h), CO/CO2/
H2/Ar = 18/10/67/5). The MARR value was calculated to be 7.58%.

3.3. Effects of operating conditions


The effects of temperature, space velocity, and pressure on the CO
conversion are provided in groups (1), (2), and (3), respectively, in
Fig. 2. The negative effects of temperature were attributed to the proximity of the conversion to equilibrium, which was negatively inuenced
by the elevated temperature in exothermic reactions [20]. However, as
shown in groups (13)(15), when the conversion was relatively far
from equilibrium, the positive effects of the temperature were observed
as a result of the reaction rates increasing as the temperature increased.
The negative effects of space velocity were attributed to the reduced
residence time, and the positive effects of the pressure were attributed
to the increased concentration of the reactants.
Alternatively, the CO2 conversions in groups (1)(3) exhibited little
variation with respect to the operating conditions, since the WGS
reaction that produces CO2 may compensate for the CO2 consumption
through hydrogenation when CO and CO2 hydrogenations occur
simultaneously (as discussed in Section 3.2, the contribution of the

146

N. Park et al. / Fuel Processing Technology 125 (2014) 139147

WGS was minimal when only CO2 hydrogenation occurred). However,


the CO fraction and the H2 fraction from the feed supply signicantly affected both the CO and CO2 conversion, as shown in groups of (4), (7),
and (8) in Fig. 2.
Fig. 3 displays the effects of temperature and CO fraction on the
MeOH concentration upon exiting the reactor, in addition to the accumulated reaction rates. As shown in Fig. 3b, the CO hydrogenation
monotonically decreased as the CO fraction decreased, while the WGS
reaction and CO2 hydrogenation exhibited maxima with respect to the
CO fraction, resulting in the maximum MeOH concentration occurring
at approximately 90% of CO fraction. Under the conditions of pressure
and space velocity described in Fig. 3 (see the caption for detailed conditions), the accumulated CO hydrogenation rates decreased as the temperature decreased because the reaction was inuenced by the reaction
rates more than the equilibrium, while both WGS and CO2 hydrogenation rates follow the thermodynamic equilibrium and increased as the
temperature decreased. Therefore, MeOH exhibited a maximum value
with respect to the temperature. It is important to note that the degree
of CO2 hydrogenation is comparable to that of CO hydrogenation when
using the commercial catalysts, while the former was approximately
one tenth of the latter when the KRICT catalysts were used [8].
3.4. Effects of particle size
Powder type catalysts were used for the kinetic data in this study;
however, the catalysts need to be shaped into a pellet in order to prevent an abrupt drop in pressure when using large scale reactors. In
such cases, the catalytic performance would deteriorate as a result of internal diffusion limitation, and additional experiments were conducted
using different particle sizes to evaluate the effects of the particle size on
the production rates. Based on the experimental data, the effectiveness
factors for each particle size were calculated [23] and regressed using an
exponent under the assumption that the factors approach zero as particle size increases. As shown in Fig. 4a, the effectiveness factors were
slightly inuenced by temperature, but substantially decreased as the
particle radius increased, and the effectiveness factor was decreased
by half when the particle radius increased 25-fold (0.1 mm to
2.5 mm). The CO conversion decreased as the particle size increased
as a result of the low effectiveness factor. The simulations conrmed
that when the particle size was small (high effectiveness factor), the
conversion was dominantly inuenced by the thermodynamic equilibrium (the higher the temperature, the lower the conversion), while a reduced effectiveness factor results in a reaction where the reaction rates
determines the conversion rate (the higher the temperature, the higher
the conversion).

regarding the effectiveness factor, it can be easily applied to the design


of large scale reactors for the synthesis of MeOH.
Nomenclature
Concentration of the specie, [mol/m3]
Ci
Heat capacity, [J/gK]
Cp
Dimensionless WeiszPrater parameter
CWP
Tube diameter, [m]
Dt
Activation energy, [J/mol]
Ej
Fugacity
fi
Objective function
Fobj
SV
Space velocity [mL/(gcat h)]
H
Heat of reaction, [J/mol]
Reaction rate constant
ki
Adsorption equilibrium constant of species i
Ki
Equilibrium constant of reaction rate
KP,i
NE
Number of experimental conditions
NR
Number of reactions
Reaction rate, [mol/(kgcat s)]
ri
R
Gas constant (R = 8.314), [J/(molK)]
s
Adsorption site
T
Temperature, [K]
Wall temperature, [K]
Tw
Gas velocity, [m/s]
us
U
Overall heat transfer coefcient, [W/m2K]
w
Weighting factor
X
Conversion [%].

Greek letters
Fraction of species i in the catalytic surface
i
Bulk pellet density, [kg/m3]
B
Bulk gas density, [kg/m3].
g

Subscripts
calc
Calculated data
i
Species
in
Inlet
j
Reaction
exp
Experimental data
ref
Reference condition.

Acknowledgment
4. Conclusions
The adsorption of CO and CO2 on various types of active sites was
assumed for the kinetic model of the synthesis of MeOH, and the corresponding reaction rates were developed using the RDS's reported in the
literature for commercial catalyst. The validity of the model with the estimated kinetic parameters was proven by comparing the experimental
and simulation data, and the model was then used to evaluate the effects of operating conditions. Analysis showed the negative effects of
space velocity due to the reduced residence time, and the positive effects of the pressure by the increased concentration of the reactants.
The developed model also showed the correlation between CO2 hydrogenation and WGS reaction, and the maximum MeOH production could
be determined by the optimal temperature and CO fraction. In addition,
the usefulness of the developed kinetic model was illustrated by showing the similar contribution of CO and CO2 hydrogenations. Additional
experiments regarding various particle sizes clearly conrmed the
change in the effectiveness factor as a result of the internal diffusion
limitation. By combining the kinetic model with the information

This work was supported by Korea Institute of Energy Technology


Evaluation and Planning (KETEP) under Energy Efciency & Resources
Programs (Project No. 2012T100201578) of the Ministry of Trade, Industry and Energy (MOTIE), Republic of Korea.
References
[1] G.A. Olah, Beyond oil and gas: the methanol economy, Angewandte Chemie International Edition 44 (2005) 26362639.
[2] B. Denise, R.P.A. Sneeden, C. Hamon, Hydrocondensation of carbon dioxide: IV,
Journal of Molecular Catalysis 17 (1982) 359366.
[3] P. Mizsey, E. Newson, T.-b. Truong, P. Hottinger, The kinetics of methanol
decomposition: a part of autothermal partial oxidation to produce hydrogen for
fuel cells, Applied Catalysis A: General 213 (2001) 233237.
[4] T. Chmielniak, M. Sciazko, Co-gasication of biomass and coal for methanol synthesis, Applied Energy 74 (2003) 393403.
[5] M.N. Mohd Fuad, M.A. Hussain, A. Zakaria, Optimization strategy for long-term
catalyst deactivation in a xed-bed reactor for methanol synthesis process,
Computers and Chemical Engineering 44 (2012) 104126.
[6] K.M.V. Bussche, G.F. Froment, A steady-state kinetic model for methanol synthesis
and the water gas shift reaction on a commercial Cu/ZnO/Al2O3 catalyst, Journal of
Catalysis 161 (1996) 110.

N. Park et al. / Fuel Processing Technology 125 (2014) 139147


[7] D.L. Chiavassa, S.E. Collins, A.L. Bonivardi, M.A. Baltans, Methanol synthesis from
CO2/H2 using Ga2O3-Pd/silica catalysts: kinetic modeling, Chemical Engineering
Journal 150 (2009) 204212.
[8] H.W. Lim, M.J. Park, S.H. Kang, H.J. Chae, J.W. Bae, K.W. Jun, Modeling of the kinetics
for methanol synthesis using Cu/ZnO/Al2O3/ZrO2 catalyst: inuence of carbon
dioxide during hydrogenation, Industrial and Engineering Chemistry Research 48
(2009) 1044810455.
[9] M. Peter, M.B. Fichtl, H. Ruland, S. Kaluza, M. Muhler, O. Hinrichsen, Detailed kinetic
modeling of methanol synthesis over a ternary copper catalyst, Chemical Engineering
Journal 203 (2012) 480491.
[10] S. Tsuchiya, T. Shiba, Adsorption measurements during the synthesis of methanol
at lower pressures, Bulletin of the Chemical Society of Japan 38 (1965)
17261729.
[11] W. von Wedel, S. Ledakowicz, W.-D. Deckwer, Kinetics of methanol synthesis in the
slurry phase, Chemical Engineering Science 43 (1988) 21692174.
[12] L.C. Grabow, M. Mavrikakis, Mechanism of methanol synthesis on Cu through CO2
and CO hydrogenation, ACS Catalysis 1 (2011) 365384.
[13] C.V. Ovesen, B.S. Clausen, J. Schitz, P. Stoltze, H. Topse, J.K. Nrskov, Kinetic
implications of dynamical changes in catalyst morphology during methanol synthesis over Cu/ZnO catalysts, Journal of Catalysis 168 (1997) 133142.
[14] K. Klier, V. Chatikavanij, R.G. Herman, G.W. Simmons, Catalytic synthesis of methanol from CO/H2: IV. The effects of carbon dioxide, Journal of Catalysis 74 (1982)
343360.
[15] M.A. McNeil, C.J. Schack, R.G. Rinker, Methanol synthesis from hydrogen, carbon
monoxide and carbon dioxide over a CuO/ZnO/Al2O3 catalyst: II. Development of a
phenomenological rate expression, Applied Catalysis 50 (1989) 265285.
[16] G. Liu, D. Willcox, M. Garland, H.H. Kung, The role of CO2 in methanol synthesis on
CuZn oxide: an isotope labeling study, Journal of Catalysis 96 (1985) 251260.
[17] C.J. Schack, M.A. McNeil, R.G. Rinker, Methanol synthesis from hydrogen, carbon
monoxide, and carbon dioxide over a CuO/ZnO/Al2O3 catalyst: I. Steady-state
kinetics experiments, Applied Catalysis 50 (1989) 247263.

147

[18] A. Coteron, A.N. Hayhurst, Kinetics of the synthesis of methanol from CO + H2 and
CO + CO2 + H2 over copper-based amorphous catalysts, Chemical Engineering
Science 49 (1994) 209221.
[19] G.H. Graaf, E.J. Stamhuis, A.A.C.M. Beenackers, Kinetics of low-pressure methanol
synthesis, Chemical Engineering Science 43 (1988) 31853195.
[20] J.M. Smith, H.C. Van Ness, M.M. Abbott, Introduction to Chemical Engineering
Thermodynamics, 7th ed. McGraw-Hill, New York, 2005.
[21] G.H. Graaf, P.J.J.M. Sijtsema, E.J. Stamhuis, G.E.H. Joosten, Chemical equilibria in
methanol synthesis, Chemical Engineering Science 41 (1986) 28832890.
[22] K.L. Ng, D. Chadwick, B.A. Toseland, Kinetics and modelling of dimethyl ether synthesis from synthesis gas, Chemical Engineering Science 54 (1999) 35873592.
[23] H.S. Fogler, Elements of Chemical Reaction Engineering, Prentice-Hall, New Jersey,
1999.
[24] H.J. Jun, M.J. Park, S.C. Baek, J.W. Bae, K.S. Ha, K.W. Jun, Kinetics modeling for the
mixed reforming of methane over Ni-CeO2/MgAl2O4 catalyst, Journal of Natural
Gas Chemistry 20 (2011) 917.
[25] M. Boudart, Two-step catalytic reactions, AIChE Journal 18 (1972) 465478.
[26] S.H. Kwack, J.W. Bae, M.J. Park, S.M. Kim, K.S. Ha, K.W. Jun, Reaction modeling on the
phosphorous-treated Ru/Co/Zr/SiO2 FischerTropsch catalyst with the estimation of
kinetic parameters and hydrocarbon distribution, Fuel 90 (2011) 13831394.
[27] N. Park, M.J. Park, S.C. Baek, K.S. Ha, Y.J. Lee, G. Kwak, H.G. Park, K.W. Jun, Modeling
and optimization of the mixed reforming of methane: maximizing CO2 utilization
for non-equilibrated reaction, Fuel 115 (2014) 357365.
[28] L.M.M. Jorge, R.M.M. Jorge, F. Fujii, R. Giudici, Evaluation of heat transfer in a catalytic
xed bed reactor at high temperatures, Braz, Journal of Chemical Engineering 16
(1999) 407420.
[29] J.H. Quinton, J. Anderson Storrow, Heat transfer to air owing through packed tubes,
Chemical Engineering Science 5 (1956) 245257.
[30] J.M. Valstar, P.J. van den Berg, J. Oyserman, Comparison between two dimensional
xed bed reactor calculations and measurements, Chemical Engineering Science
30 (1975) 723728

You might also like