You are on page 1of 43

Subscriber access provided by Fudan University

Article
The Interplay of Kinetics and Thermodynamics in the
Catalytic Steam Methane Reforming over Ni/MgO-SiO2
Naoki Kageyama, Brigitte R. Devocht, Atsushi Takagaki, Kenneth
Toch, Joris W. Thybaut, Guy B Marin, and S. Ted Oyama
Ind. Eng. Chem. Res., Just Accepted Manuscript DOI: 10.1021/acs.iecr.6b03614 Publication Date (Web): 10 Jan 2017
Downloaded from http://pubs.acs.org on January 11, 2017

Just Accepted

Just Accepted manuscripts have been peer-reviewed and accepted for publication. They are posted
online prior to technical editing, formatting for publication and author proofing. The American Chemical
Society provides Just Accepted as a free service to the research community to expedite the
dissemination of scientific material as soon as possible after acceptance. Just Accepted manuscripts
appear in full in PDF format accompanied by an HTML abstract. Just Accepted manuscripts have been
fully peer reviewed, but should not be considered the official version of record. They are accessible to all
readers and citable by the Digital Object Identifier (DOI). Just Accepted is an optional service offered
to authors. Therefore, the Just Accepted Web site may not include all articles that will be published
in the journal. After a manuscript is technically edited and formatted, it will be removed from the Just
Accepted Web site and published as an ASAP article. Note that technical editing may introduce minor
changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers
and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors
or consequences arising from the use of information contained in these Just Accepted manuscripts.

Industrial & Engineering Chemistry Research is published by the American Chemical


Society. 1155 Sixteenth Street N.W., Washington, DC 20036
Published by American Chemical Society. Copyright American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
The Interplay of Kinetics and Thermodynamics in the Catalytic Steam Methane Reforming over
5
6
7
8
Ni/MgO-SiO2
9
10
11 Naoki Kageyamaa, Brigitte R. Devochtb, Atsushi Takagakia,c, Kenneth Tochb, Joris W. Thybautb*, Guy
12
13
14 B. Marinb, S. Ted Oyamaa,c,d*
15
16 a
17 Department of Chemical System Engineering, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku,
18
19
20 Tokyo, 113-8656, Japan
21
22
b
23 Laboratory for Chemical Technology, Ghent University, Tech Lane Ghent Science Park Campus A,
24
25
26 Technologiepark 914, B-9052 Ghent, Belgium
27
28
29 c
College of Chemical Engineering, Fuzhou University, Fuzhou 350116, China
30
31
32 d
Environmental Catalysis and Nanomaterials Laboratory, Department of Chemical Engineering,
33
34
35 Virginia Polytechnic Institute and State University, Blacksburg, VA 24061-0211, United States
36
37
38
39
40
41 *Corresponding authors: STO oyama@vt.edu, JWT joris.thybaut@ugent.be
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 1
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 2 of 42

1
2
3
4 ABSTRACT
5
6
7
The steam methane reforming (SMR) reaction was studied on a Ni/MgO-SiO2 catalyst at 923 K (650
8 o
9 C) and 0.40 MPa in a tubular packed-bed reactor. The partial pressures of CH4 and H2O were varied
10
11 between 20 and 140 kPa and 80 and 320 kPa respectively. Measurements were carried out without mass
12
13 and heat transport limitations, as verified by the Weisz-Prater and Mears criteria. Experimentally, the
14
15
CH4 conversion increased with the inlet partial pressure of H2O and decreased with the inlet partial
16
17
18 pressure of CH4. However, at low CH4 inlet partial pressures, i.e., at 40 and 60 kPa, the conversion
19
20 passed through a maximum. Rate expressions were derived based on a simple two-step sequence. A
21
22 statistical analysis led to a globally significant, weighted regression and resulted in a good agreement
23
24 between the model and the experimental data, as indicated by a low F value of model adequacy of 2.84.
25
26
27 The rate and equilibrium coefficient parameters were statistically significant as indicated by narrow
28
29 confidence intervals. The model was able to correctly describe the experimentally observed maximum
30
31 in the methane conversion and allowed to relate this behavior to CH4 and H2O surface coverages. The
32
33 model was able to capture the increasing selectivity of CO2 with the H2O inlet partial pressure and
34
35
36
increasing methane conversion. The effect of changing the total pressure and H2O/CH4 ratio on the CH4
37
38 conversion as a function of the space velocity was simulated and corresponded to both the experimental
39
40 and literature data. A major finding of the modelling is that as flow rate is increased there is a
41
42 crossover in the order of conversion with pressure due to a transition from thermodynamic to kinetic
43
44 control. Although the SMR equilibrium conversion decreases with pressure, away from equilibrium at
45
46
47 high flow rates, conversion is higher at higher pressures due to enhanced adsorption rates.
48
49 Keywords: steam methane reforming; Ni/MgO-SiO2 catalyst; integral kinetic analysis; kinetics and
50
51 thermodynamics
52
53
54
55
56
57
58
59 2
60
ACS Paragon Plus Environment
Page 3 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
1. INTRODUCTION
5
6
7
8
9
Hydrogen plays an important role in todays chemical industry, especially in the refining of petroleum
10
11
12 and the synthesis of chemical products. Recently, hydrogen has received attention as a potential energy
13
14
15 source for fuel cells because of its high energy conversion efficiency 1. The majority of hydrogen is
16
17
18 produced by fossil fuel reforming 2 , especially via steam reforming of natural gas, whose main
19
20
21 component is methane. Steam Methane Reforming (SMR) is a strongly endothermic reaction (Eq. 1)
22
23
24 which is generally accompanied by the moderately exothermic water-gas shift reaction (Eq. 2).
25
26
27
28 Ho298 = +206 kJ mol-1 Eq. 1
29
30
31 Ho298 = -41 kJ mol-1 Eq. 2
32
33
34 Both reactions contribute to the production of H2. In contrast, the much studied dry-reforming of
35
36
37
38
methane (CH4 + CO2 2 CO + 2H2) cannot be used for commercial hydrogen production
39
40
41 because the reverse water-gas shift reaction (CO2 + H2 CO + H2O) consumes hydrogen to
42
43
44
produce water. 3 The SMR process is operated industrially at 0.3-2.5 MPa total pressure and at high
45
46
47 temperature, i.e., typically between 900 and 1300 K. The energy requirement is considerable and is
48
49
50 commonly supplied by the combustion of natural gas, c.q., methane, in a furnace in which the SMR
51
52
53 reactor is placed. In general, this results in a high energy consumption and large CO2 emissions. The
54
55
56 simultaneously occurring water-gas shift reaction 4 produces additional H2 and CO2. There have been
57
58
59 3
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 4 of 42

1
2
3
4
many attempts to improve the process efficiency. In a study by Zheng et al., the energy required for the
5
6
7
8
endothermic reforming was supplied using solar energy as a renewable energy source 5 . In an
9
10
11 investigation by Hafizi et al., chemical looping was employed by cycling fine metal powders between
12
13
14 oxidation and reduction steps for hydrogen production from methane 6. In addition to these attempts,
15
16
17 efforts have been made to develop new catalysts.
18
19
20 Kinetic modeling bridges the gap between phenomena occurring at the conceptual and practical scale.
21
22
23 It provides essential information for reaction design and corresponding process scale-up 7 as well as for
24
25
26 understanding catalyst deactivation 8. It can also be used to better understand the catalyst activity so as
27
28
29 to allow its improvement. As such a better understanding is pursued in this work with a model
30
31
32 accounting for the most kinetically relevant steps. Although some studies aim for increasingly more
33
34
35 detailed descriptions of the reaction 9,10,11 this work takes an opposite approach to produce the simplest
36
37
38 possible analysis. The significance of the work is two-fold. First, it presents a minimal kinetic model
39
40
41
with just two equilibrium parameters and two rate parameters to describe the MSR and WGS reactions
42
43
44
45
that can be used by other researchers as a metric for comparison. Second, it demonstrates that the
46
47
48 conversion falls with increasing pressure at low space velocity (high space time) as expected from
49
50
51 thermodynamics, but increases with increasing pressure at high space velocity (low space time) from
52
53
54 kinetic control. As far as we know, this interplay between thermodynamics and kinetics mediated by
55
56
57 pressure has not been noted before.
58
59 4
60
ACS Paragon Plus Environment
Page 5 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
The catalyst employed in this research is Ni-based, which is the most commonly used metal, but
9
10
11 employs a MgO-SiO2 support that has not been studied widely. Magnesium silicate can be made in
12
13
14 high surface area and has been used as a catalyst [12] as well as a catalyst support [13]. It is strong,
15
16
17 low cost, and offers sulfur tolerance, which is a desirable property with streams derived from natural
18
19
20 gas. In addition, Ni on this support is easier to reduce than on Al2O3-based supports because
21
22
23 aluminates are not formed. In this research a series of steam methane reforming measurements was
24
25
26 conducted by varying the ratio of the inlet gas components, methane and water vapor as reactants and
27
28
29 nitrogen as inert dilution gas at a pressure of 0.40 MPa.
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 5
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 6 of 42

1
2
3
4
2. EXPERIMENTAL SECTION
5
6
7
8
2.1 Experimental
9
10
11 The gases used in this study were CH4 (99 % Toatsu Yamazaki Co., Ltd.), H2 (99.99 % Showa Denko
12
13
14 Gas Products Co., Ltd.) and N2 (99.99 % Toatsu Yamazaki Co., Ltd.). A conventional Ni/MgO-SiO2
15
16
17 catalyst (JGC Catalysts and Chemicals Ltd.) consisting of 53.0-58.0 wt% NiO, 8.1-11.1 wt% MgO and
18
19
20 24.2-28.2 wt% SiO2 (supplier information) was selected. This commercial catalyst is used for naphtha
21
22
23 reforming and is, hence, chemically sufficient stable enough at high pressure and temperature. The BET
24
25
26 area was determined by N2 adsorption at 77 K using a volumetric apparatus (BELSORP-mini, Microtrac
27
28
29 BEL Corp.). The H2 chemisorption uptake was measured using a pulse flow instrument (CHEMBET-
30
31
32 3000, Quantachrome Instruments) at 313 K after 12 h reduction at 923 K in 5 % H2/Ar.
33
34
35
36 Steam methane reforming was studied in a vertical concentric tubular reactor of a geometry suitable for
37
38
39 future membrane reactors studies. The reactor consisted of an inner dense alumina tube, a quartz sleeve
40
41
42
43
and an outer stainless steel shell and was heated by an external electric furnace, see Fig.1. An amount
44
45
46 of 0.1 g of Ni/MgO-SiO2 catalyst diluted with 3.9 g SiO2 particles (Wako Pure Chemical Industries
47
48
49 Ltd.) formed the catalyst bed of volume 3.0 cm3. For optimal heat transfer, the catalyst bed was placed
50
51
52 in the outer annular region between the quartz tube and the dense alumina tube. It was held in the middle
53
54
55 of the reactor by extra SiO2 particles filling the lower half of the reactor, see Fig. 1. The outer tubular
56
57
58
59 6
60
ACS Paragon Plus Environment
Page 7 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
stainless steel tube covered the quartz tube for mechanical strength and rigid sealing, but an opening
5
6
7
8
allowed pressure equalization between its inside and the outside of the quartz sleeve. The reactor
9
10
11 pressure was controlled by a backpressure regulator connected at the reactor outlet and the reactor
12
13
14 temperature was monitored by a thermocouple placed at the bottom of the catalyst bed. A total pressure
15
16
17 of 0.40 MPa and a temperature of 923 K was used for the measurements. The catalyst particles were
18
19
20 sieved to 400-630 m in diameter and the SiO2 particles used for dilution of the catalyst to 600-850 m.
21
22
23 Upstream of the reactor, CH4 and N2 were mixed and introduced to a vaporizer heated at 473 K. Liquid
24
25
26 water was fed using a pump (Hitachi L-7100) to the vaporizer and immediately vaporized and mixed
27
28
29 with the N2/CH4 flow. The mixed gas flowed to the reactor in a heated tube to prevent water
30
31
32 condensation in the transfer lines. The inlet gas composition was varied by keeping the total flow rate
33
34
35 at 270 mol s-1 (400 cm3 (NTP) min-1), corresponding to a constant gas-hourly space velocity of 2000
36
37
38 h-1. The weight-hourly space velocity (WHSV) and space time (W/F0CH4) was 12000-72000 Ncm3 gcat-
39
40
41 1
h-1 and 1.12 - 6.72 kgcat s mol-1 respectively with respect to CH4. The inlet flow composition varied in
42
43
44
45
the range of 5-35 vol% for CH4 and 20-70 vol% for H2O, diluted with nitrogen. The space time is
46
47
48 inversely proportional to the weight-hourly space velocity, see Eq. 3, with Tn and pn the normal
49
50
51 temperature and pressure of 273.15 K and 101325 Pa, W the catalyst mass [kg], F0CH4 the inlet flow rate
52
53
54 of methane [mol s-1] and R the universal gas constant [J mol-1 K-1].
55
56
57
58
59 7
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 8 of 42

1
2
3

4
0 = Eq. 3
5 4

6
7
8
9
The outlet gas flow was analyzed by a gas chromatograph (GC-8A, Shimadzu Co.) equipped with a
10
11
12 thermal conductivity detector (TCD) using a combination of Porapak-Q, Shincarbon and Molecular
13
14
15 sieve 5A columns. However, it was difficult to accurately detect the precise amount of H2O with the
16
17
18 TCD due to its large amount and the high affinity between H2O and the GC columns. The amount of
19
20
21 H2O was calculated from the mass-balance averages of hydrogen and oxygen individually. For
22
23
24 interpreting the dataset, the experimental data was normalized by closing the carbon balance, ensuring
25
26
27 that the total carbon flow rate at the reactor inlet equals that at the reactor outlet.
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 Fig. 1. Reactor used in the steam methane reforming studies
54
55
56
57
58
59 8
60
ACS Paragon Plus Environment
Page 9 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
2.2 Intrinsic kinetics regime verification
5
6
7
8
9
Prior to the kinetic modelling, the Weisz-Prater and the Mears criteria were calculated to assess the
10
11
12 absence of mass and heat transfer limitations 14. The Mears criterion,
13
14
15
16 < 0.15 Eq. 4
17
18
19
20 was calculated to be equal to 0.049 at the reaction condition which gave the highest CH4 conversion
21
22
23 58 % and the Weisz-Prater criterion,
24
25
26 2
4 cat
27 < 0.3 Eq. 5
28 4e CH4 ,b
29
30
31 was calculated as 0.032 at the same reaction conditions. It can, hence, be concluded that there were no
32
33
34 internal nor external heat or mass transfer limitations in the present study.
35
36
37
38 Definitions and actual values for the quantities in Eqs. 4 and 5 are given in Table 1 below. The highest
39
40
41
CH4 conversion of 58% was obtained with 40 kPa CH4, 200 kPa H2O and 160 kPa N2., W = 100 mg,
42
43
44
45
vtot = 400 cm3 (NTP) min-1.
46
47
48
49
50
51
52
53
54
55
56
57
58
59 9
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 10 of 42

1
2
3
4
5 Table 1. Quantities in the Mears and Weisz-Prater Criteria
6
7 Quantity Definition Values Units
8
9
10
H Heat of the reaction of SMR -2.04x105 J mol-1
11
12 RCH4 Net rate of formation of CH4 0.169 mol kgcat-1 s-1
13
14 d Specific length of the reactor 2.5x10-3 m
15
16
17 h Heat transfer coefficient 770 J m-2 K-1 s-1
18
19 T Temperature 923 K
20
21 Ea Activation energy of SMR 1.5x105 J mol-1
22
23
24 R Gas constant 8.31 J mol-1 K-1
25
26 cat Catalyst density 5180 kg m-3
27
28 dp Catalyst pellet diameter 5.16x10-4 m
29
30
31 n Reaction order 1
32
33 kc Mass transfer coefficient 0.173* m s-1
34
35
CCH4,b Bulk gas concentration of CH4 5.51 mol m-3
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 10
60
ACS Paragon Plus Environment
Page 11 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
Parameter estimation
5
6
7
8
9
The Iteratively Reweighted Least Squares (IRLS) method has been used as regression procedure for model
10
11
12 parameter estimation 15. This procedure consists of minimizing the sum of squares of the residuals e, i.e.,
13
14
15 the difference between the experimental observations and model simulations, which approximate the
16
17
18 experimental error with respect to the model parameters in which the variance-covariance matrix D
19
20
21 is taken into account, see Eq. 6. This matrix is required because of the heteroscadisticity of the
22
23
24 experimental error, i.e., its variance depends on the operating conditions and weighs each of the elements
25
26
27 within the sum of squares in a statistically justified manner.
28
29
S ( ) = D ( ) Eq. 6
1
Min
T
30
31
32
33 The response weights, i.e., the elements of the inverse of the variance-covariance matrix are automatically
34
35
36 determined within the modelling software used, i.e., the MicroKinetic Engine, and updated after each
37
38
39 regression until convergence in the weights was achieved. Typically, a few iterations sufficed to reach this
40
41
42
43
convergence.
44
45
46 The approximation of the true model parameters with the most optimal parameter estimates b yields the
47
48
49 final residuals e resulting in the minimization of the objective function S, see Eq. 7.
50
51
52 = y X
e = y X b Eq. 7
53
54
55
56
57
58
59 11
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 12 of 42

1
2
3
4
The tubular reactor is simulated as an ideal plug flow reactor, operated in steady state. The set of continuity
5
6
7
8
equations for all components forms the reactor model, shown by Eq. 8 with Ri the net rate of formation of
9
10
11 component i and W the catalyst mass.
12
13
14 dFi = Ri dW Eq. 8
15
16
17 The independent variables that serve as an input for the model during the regression procedure are the
18
19
20 space time, the temperature, the total pressure and the inlet molar flow rate of the five components, i.e.,
21
22
23 methane, water, hydrogen, carbon monoxide and carbon dioxide. The five model responses are the
24
25
26 corresponding outlet molar flow rates of the same components. In total, 37 experiments, including 14
27
28
29 repetition experiments, were taken into account for the regression.
30
31
32 A model is deemed to be adequate if the difference between the experimentally observed responses
33
34
35 and the model calculated responses can be mainly attributed to the experimental error 16. The model
36
37
38 adequacy is assessed via an F test, in which an Fa value is defined as the ratio of the lack-of-fit sum of
39
40
41
squares and the pure-error sum of squares (Eq. 9). The former is calculated as the difference between the
42
43
44
45
residual sum of squares and the latter as the pure-error sum of squares of the variance of repeated
46
47
48 observations 17. The test is based on the null hypothesis that the model is adequate and states that Fa is
49
50
51 distributed to F with dfLOF degrees of freedom in the numerator and dfPE degrees of freedom in the
52
53
54 denominator. If the calculated F value is smaller than the corresponding tabulated F value for the selected
55
56
57 confidence level, the null hypothesis is accepted and the model is considered to be adequate. In this work
58
59 12
60
ACS Paragon Plus Environment
Page 13 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
a confidence level of 0.95 is selected, corresponding with a tabulated F value in the range of 2 5, the
5
6
7
8
exact value being determined by dfLOF and dfPE. It should be noted that the adequacy test is a severe test
9
10
11 which is rarely fulfilled.
12
13 SSQ LOF
14 df LOF
15 Fa =
16 SSQ PE
17 df PE
18 Eq. 9
19
20 In order to assess the global significance of the regression, the hypothesis that all parameters would be
21
22
23 equal to zero simultaneously is verified, via an F test. By means of this test it is determined whether the
24
25
26 actual model performs better than a model which takes the average of the observed experimental values
27
28
29 as response values. The Fs value is calculated as the ratio of the regression sum of squares and the residual
30
31
32 sum of squares (Eq. 10). If the calculated F value exceeds the tabulated at the selected confidence level,
33
34
35 the null hypothesis is rejected and the regression is seen as meaningful and globally significant. In practice,
36
37
38 the null hypothesis is easily rejected and the calculated F values should at least be of the order of magnitude
39
40
41
of 100.
42
43
44 SSQ REG
45 df REG
46 Fs =
47
SSQ RES
48 df RES Eq. 10
49
50
51
52
53
54
55
56
57
58
59 13
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 14 of 42

1
2
3
4
5
6
7
8
9
10
11 2.4 Software
12
13
14 The model in this work was regressed to the experimental data and analyzed with the MicroKinetic Engine
15
16
17 (KE) 18. The KE is a software tool for the simulation and regression of chemical kinetics and has been
18
19
20 developed at the Laboratory for Chemical Technology at Ghent University. Two deterministic regression
21
22
23 routines are implemented in the software, i.e., the Rosenbrock and Levenberg-Marquardt algorithm which
24
25
26 are consecutively used during the process. The KE is mainly characterized by its user-friendliness,
27
28
29 robustness and flexibility.
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 14
60
ACS Paragon Plus Environment
Page 15 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
3. RESULTS AND DISCUSSION
5
6
7
8
9
3.1 State-of-the-art and catalyst choice
10
11
12
13 This study employed a Ni/MgO-SiO2 catalyst. The BET surface area was 105 m2g-1 and the hydrogen
14
15
16 uptake, Vads, was 90 mol g-1. The hydrogen uptake (Vads) can be used to calculate the particle size 19
17
18
19 from the formula, d = 6c/VadswNA where c is the supported Ni metal weight % (43.2wt%), is the
20
21
22 bulk Ni density (8.91 g cm-3), is the surface Ni atom density (0.0649 nm2/atom), w is the weight of
23
24
25 catalyst, and NA is Avogadros number. The calculated particle size was 41 nm. The dispersion of the
26
27
28 catalyst can be obtained from the hydrogen uptake and the Ni content of the catalyst, D = VadsMWNi/c,
29
30
31 and is found to be 1.2%, which is consistent with the large particle size. Measurements on the catalyst
32
33
34 were carried out for more than two weeks, while verifying reproducibility of points. This sets the
35
36
37 stability of the catalyst at least 200 h. Fresh catalyst was loaded as needed for different measurements.
38
39
40
41
Recent developments in the area of catalytic SMR were described in a 2013 review 20. Here, the latest
42
43
44
45
catalysis studies since then are summarized. Using conventional SMR reactors 21, most of the papers
46
47
48 report Ni as the principal component but use different supports such as CaO-Ca5Al6O14 22, -Al2O323,
49
50
51 K2TixOy-Al2O3 24, -Al2O3 25, NiAl2O4 26, TiO2 27, SiO2 28, ZrO2 29,Al2O3 (Ca co-loaded as Ca-Ni/Al2O3)30,
52
53
54 SBA-15 31. Ni has been reported to change the distribution of certain co-loaded metals and improve their
55
56
57 catalytic activity for other reactions too, apart from SMR 32 . The performance of those catalysts is
58
59 15
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 16 of 42

1
2
3
4
summarized in Table 2 at reaction conditions that are similar to those used in the present investigation.
5
6
7
8
As can be seen, some of the experimental results are at or close to equilibrium. The present studies are
9
10
11 carried out away from equilibrium in order to obtain data in the kinetic regime.
12
13
14
15
16 Table 2. Summary of recent results in steam methane reforming
17
18
19 WHSV (space
20
21 time)
T P XCH4 Equil.XCH4
22 Catalyst Ni content H2O/CH4 Ncm3CH4 h-1 CO2/CO Stability Ref.
23 K MPa % %
24 gcat-1 (kgcat s
25 molCH4-1)
26
27 Ni/CaO-Ca5Al6O14 15 wt% Ni 923 0.10 4 230 (350.83) 95 95 0.48 NR 22
28
29 Ni/-Al2O3 20 wt% Ni 928 0.10 3 15000 (5.38) 88 98 NR NR 23
30
31 Ni/K2TixOy-Al2O3 11 wt% Ni 973 0.10 2.5 15000 (5.38) 86 96 NR >100 h 24
32
33 Ni/-Al2O3 20 wt% Ni 973 0.10 3 35000 (2.31) 84 97 1.6 NR 25
34
35
NiAl2O4
36 33 wt% Ni 923 0.10 3 38000 (2.12) 80 96 0.70 NR 26
37 (spinel structure)
38
39 Ni/TiO2 10 wt% Ni 773 0.10 1 6000 (13.45) 27 29 2.4 > 96 h 27
40
41
10 wt%
42 Ni/SiO2 973 0.10 0.5 12000 (6.72) 20 83 NR NR 28
43 NiO
44
45 Ni/ZrO2 15 wt% Ni 873 0.10 1 39000 (2.07) 20 71 1.2 4h 29
46
47 64 wt%
48 Ca-Ni/Al2O3 873 0.10 4 NR 85 95 0.36 NR 30
49 NiO
50
51 Ni/SBA-15 25 wt% Ni 823 0.10 2 NR 21 48 NR NR 31
52
53 56 wt%
54 Ni/MgO-SiO2 923 0.40 2 48000 (1.68) 40 61 2.1 > 200 h *
NiO
55
56
57 NR: Not reported, * This work
58
59 16
60
ACS Paragon Plus Environment
Page 17 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
A more extended range of operating conditions has been investigated as part of the present work. Of
5
6
7
8
course, some repetition experiments have been performed, i.e., 10 experiments with feeds consisting of
9
10
11 CH4 20 %, H2O 40 %, N2 40 %, 2 experiments with CH4 20 %, H2O 60 %, N2 20 % and 2 experiments
12
13
14 with CH4 25 %, H2O 60 %, N2 15 %. The relationship between the methane conversion and the WHSV
15
16
17 is plotted in Fig. 2 for the studies reported in Table 1. As can be seen, the data can be categorized in two
18
19
20 regions, each exhibiting a decreasing conversion with increasing WHSV. The measurements with a
21
22
23 H2O/CH4 ratio exceeding 2 resulted in a higher CH4 conversion than those with a H2O/CH4 ratio below
24
25
26 2. Variation of the points from the trend lines is expected as there is some variation in temperature and
27
28
29 H2O/CH4 ratio, even though the reported studies all used a pressure of 0.10 MPa. The results obtained
30
31
32 as part of the present work with a H2O/CH4 ratio equal to 2 and with a higher pressure of 0.40 MPa are
33
34
35 situated in the middle of the two regions. The filled square shows a simulated result, which will be
36
37
38 discussed later, see section 3.
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 17
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 18 of 42

1
2
3
4 Space time / kgcat s mol-1CH
5 4

6 -- 8.07 4.03 2.69 2.02 1.61


7 100
8 (20)
(19) (22)
9
10 80 (21)
(23)
CH4 conversion / %
11
12 H2O/CH4 = 2.5 ~ 4
13 60
14 This work
15 H2O/CH4 = 2
16
17 40
18
19 H2O/CH4 = 0.5 ~ 1
20 20 (19)
21 (25) (26)
22
23
24
0
25
0 10000 20000 30000 40000 50000
3 -1 -1
26 WHSV / Ncm CH4
h g cat
27
28
29
Fig. 2. Relationship between the conversion of methane and WHSV/space time in recent studies on
30
31 steam methane reforming with Ni-based catalysts. Reference numbers are indicated by each point and
32
33 correspond to Table 1. The lines are a guide for the eye. : H2O/CH4 > 2, : H2O/CH4 < 2, : this
34
35 work (H2O/CH4/N2 = 2/1/2, total inlet flow 400 Nml min-1, 0.4 MPa, 923 K), : simulated result
36
37 (H2O/CH4/N2 = 2/1/2, total inlet flow 400 Nml min-1, 0.10 MPa, 923 K), by solving the set of Eq. 8
38 with the net rates of formation given by Eqs. 19 to 23. The rate and equilibrium coefficients are
39
40 estimated by a weighted regression and are shown in Table 2.
41
42
43
44 3.2 Reaction description and model construction
45
46
47
48 In this study, SMR is described with a simple model consisting of four steps, which are defined as (1)
49
50
51 CH4 adsorption on the catalyst surface, (2) H2O adsorption on the catalyst surface, (3) the oxidation of
52
53
54 CH4 towards CO and H2 and (4) the water-gas shift reaction. This combination of steps was chosen
55
56
57 as the simplest possible set from a kinetic standpoint, and is not meant to represent an actual mechanism
58
59 18
60
ACS Paragon Plus Environment
with elementary steps. In fact it is unlikely for methane and water to adsorb molecularly, and the

adsorbed species shown are just representations of species such CHx, C, CO, or H As such the

parameters of the reaction will be called equilibrium coefficients and rate coefficients, rather than

equilibrium constants and rate constants.

Simple Model

Reaction (1)

Reaction (2)

Reaction (3)

Reaction (4)

We have also considered a case where Reaction (4), the water-gas shift reaction, is reversible. This

will be discussed below and the results presented in the Supplementary Information.

The asterisk indicates adsorbed species or empty sites. In this model the products CO, CO2 and H2 are

assumed to desorb from the catalyst instantaneously, i.e., no adsorption is taken into account for the

products. The model combines a surface reaction for the reforming step with a gas-surface reaction

for the water-gas shift reaction. The latter does not imply a mechanism and is formulated simply to
19
Industrial & Engineering Chemistry Research Page 20 of 42

1
2
3
4
avoid an explicit term containing adsorbed CO.
5
6
7
8
9
Aside from the equilibrium adsorption steps, the other steps are considered to be irreversible, with the
10
11
12 reverse reactions not being involved. This is a simplification, made in order to reduce the number of
13
14
15 fitting parameters in this model. Since much of the data are at relatively low conversion, this is
16
17
18 permissible. The effect of the reverse reaction is incorporated into the reaction rate parameters, which
19
20
21 will have lower values than if the back reaction were explicitly considered. This will be discussed
22
23
24 later.
25
26
27
28 The equilibrium coefficients for Reaction (1) is K1 and for Reaction (2) is K2. Eq. 11 and 12 show the
29
30
31 net reaction rate of step (3) and (4) with k3 and k4 the corresponding rate coefficients. The reactant
32
33
34 concentrations are expressed in terms of surface concentrations because, contrary to the products, CH4
35
36
37 and H2O react from the adsorbed state.
38
39
40 r3 = k 3CCH * C H' O* Eq. 11
41 4 2

42
43 r4 = k 4 pCO C H O* Eq. 12
44 2

45
46
47 When two surface species are involved in a reaction step as in Reaction (3), the surface species cannot
48
49
50 occupy just any sites, but these sites need to be next to each other for the reaction to take place. This
51
52
53 adjacent site requirement has been treated in a textbook 33, and is accounted for in the rate expression
54
55
56 of the reaction step, using the number of adjacent neighbors z and Ctot the total concentration of
57
58
59 20
60
ACS Paragon Plus Environment
Page 21 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
active sites as shown in Eqs. 13 and 14. When properly accounting for the dual site mechanism in the
5
6
7
8
rate expression of the reforming step, the reaction rate varies linearly with the total concentration of
9
10
11 active sites, and not quadratically. 34
12
13
14 The new rate expressions r3 and r4 , shown in Eq. 14 and Eq. 15, are obtained by writing the
15
16
17 concentration of the surface species in terms of the surface coverage 35.
18
19
20 C H O*
21 C H' O* = z 2
= z H 2O Eq. 13
2
Ctot
22
23
24 r3 = k 3 zCCH * H 2O = k 3 zCtot CH 4 H 2O Eq. 14
25 4

26
27
28
r4 = k 4 pCO Ctot H 2O Eq. 15
29
30
31 After substituting the surface coverages with observable quantities, i.e., the partial pressures of the
32
33
34 components, the final expression of the reaction rate equations is obtained. The site balance (Eq. 16)
35
36
37 and the expressions for quasi equilibrium of the adsorption/desorption steps (Eq. 17 and 18), allow to
38
39
40 write the rate expression in terms of the partial pressure of methane, water and carbon monoxide (Eq.
41
42
43 19 and 20).
44
45
46 1 = * + CH 4 + H 2O Eq. 16
47
48
49 k1+
50 CH = pCH 4 * = K1 pCH 4 * Eq. 17
51
4
k1
52
53
k 2+
54 H O = p H O * = K 2 p H 2O * Eq. 18
55 2
k 2 2
56
57
58
59 21
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 22 of 42

1
2
3
k 3 K 1 K 2 zC tot p CH 4 p H 2O k 3' K 1 K 2 p CH 4 p H 2O
4 r3 = =
5 (1 + K 1 p CH 4 + K 2 p H 2O )2
(1 + K 1 p CH 4 + K 2 p H 2O )
2 Eq. 19
6
7 k 4 K 2 C tot p CO p H 2O k 4' K 2 p CO p H 2O
8 r4 = = Eq. 20
9 1 + K 1 p CH 4 + K 2 p H 2O 1 + K 1 pCH 4 + K 2 p H 2O
10
11
12 In order to solve the plug flow reactor model, see Eq. 8, the rate expressions are substituted in the net
13
14
15 rate of formation Ri for each component (Eqs. 21 to 5).
16
17
18 RCH 4 = r3 Eq. 21
19
20
21 RH 2 O = r3 r4 Eq. 22
22
23
24 RCO = r3 r4 Eq. 23
25
26
27 RCO2 = r4
28 Eq. 24
29
30
RH 2 = 3r3 + r4 Eq. 25
31
32
33
34 3.3 Model and parameter estimates assessment and discussion
35
36
37 Fig. 3a,b) show the same data of CH4 conversion as a function of the H2O inlet partial pressure for CH4
38
39
40 inlet partial pressures amounting to 40 kPa, 80 kPa, and 120 kPa. The lines represent the results of the
41
42
43 weighted regression. As expected, the CH4 conversion was higher when its inlet partial pressure was lower.
44
45
46 The CH4 conversion increased with increasing H2O partial pressure, although at high H2O inlet partial
47
48
49 pressure the CH4 conversion tended to decline. This behavior can be explained by the CH4 inhibition on
50
51
52 the catalyst surface by adsorbed H2O.
53
54
55
56 The numbers in Fig. 3a represent the equilibrium quotients for the MSR reaction, defined as
57
58
59 22
60
ACS Paragon Plus Environment
Page 23 of 42 Industrial & Engineering Chemistry Research

1
2
3
4 2 3 1
= 2
Eq. 26
5 4 2
6
7
8
At equilibrium at 923 K, QSMR is equal to the equilibrium constant, K = 2.76. All the points are in a
9
10
11 state far away from equilibrium.
12
13
14 The numbers in Fig. 3b represent the equilibrium quotients for the WGS reaction, defined as
15
2 2
. Eq. 27
16
17 =
2
18
19
20 At equilibrium QWGS is equal to the equilibrium constant, K= 2.04. No point was in the equilibrium
21
22
23 state, but some were relatively close.
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 23
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 24 of 42

1
2
3
4 65
5 0.0053 0.0033
a) 0.0054
6 0.0083

CH4 conversion / %
7 55
8 0.012
9 0.0006 0.0004
10
11 45
12 Numbers:
13 QSMR 0.0093 0.0067
14 35 0.015 0.012
15
0.011
16
0.013
0.011
17 25
18 40 80 120 160 200 240 280 320 360
19
20 H2O inlet partial pressure / kPa
21
22
23 65
1.17
24 b) 1.31 1.21 Value of QWGS
25 1.24
CH4 conversion / %

26 55
27
1.15
28 1.56 1.37
29 45
30
31 1.43
32 0.96 1.17
35 1.16
33
34 1.06 1.02
35 1.06
36 25
37 40 80 120 160 200 240 280 320 360
38 H2O inlet partial pressure / kPa
39
40
41 Fig. 3. Simple model results for CH4 conversion as a function of the H2O inlet pressure. Points:
42 experimentally observed, lines: calculated by solving the set of Eq. 8 with the net rates of formation
43
44 given by Eqs. 19 to 23. Results are for at a total pressure of 0.40 MPa and a temperature of 923 K. The
45 rate and equilibrium coefficients are obtained by a weighted regression and are shown in Table 2. , full
46
47 line, black - 40 kPa CH4 inlet partial pressure (p0CH4) or a space time of 3.36 kg s mol-1, , dashed line,
48 red - 80 kPa p0CH4 or 1.68 kg s mol-1, , dotted line, blue - 120 kPa p0CH4 or 1.12 kg s mol-1.
49
50 a) The numbers shown are the equilibrium quotients for the methane steam reforming, QSMR, with a
51
52
value at equilibrium of 2.76. Conditions are far from equilibrium.
53 b) The same data are shown but now the numbers show the equilibrium quotients for the water-gas shift
54
55 reaction, QWGS, with a value at equilibrium of 2.04. Conditions are closer to equilibrium.
56
57
58
59 24
60
ACS Paragon Plus Environment
Page 25 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
Because the data for the WGS reaction were relatively close to equilibrium, calculations were carried
5
6
7 out to ascertain whether the reverse reaction could influence the results of the fitting of the present data.
8
9 Details are reported in the Supplementary Information. The analysis indicates that there was no
10
11
12 significant influence on the results. Calculated values for the kinetic parameters are shown in Table 3,
13
14 to be discussed later.
15
16
17 Fig. 4 shows the selectivity to CO and CO2 as a function of the H2O inlet partial pressure for a CH4 inlet
18
19 partial pressure amounting to 80 kPa. The model is able to simulate the correct trend for the selectivities
20
21
22 with varying H2O partial pressure. Nevertheless there is a systematic deviation between the experimentally
23
24 observed and calculated data.
25
26
27 35 85
28
29
30

Selectivity to CO2 / %
30 80
Selectivity to CO / %

31
32
33
34
35 25 75
36
37
38
39
40 20 70
41
42
43
44 15 65
45 160 200 240 280
46
47
H2O inlet partial pressure / kPa
48
49 Fig. 4. Selectivity to CO and CO2 as a function of the inlet partial pressure of H2O at a total pressure of
50
51 0.4 MPa, a temperature of 923 K, a space time of 1.44 kg s mol-1 and conversions between 40 % and 44 %.
52 Points: experimentally observed, lines: calculated by solving the set of Eq. 8 with the net rates of formation
53
54 given by Eqs. 19 to 23. The rate and equilibrium coefficients are estimated by a weighted regression and
55 are shown in Table 2. , full line, black - selectivity to CO, , dashed line, red - selectivity to CO2.
56
57
58
59 25
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 26 of 42

1
2
3
4
5
6
7 Fig. 5 shows the CO and CO2 selectivity as a function of CH4 conversion at a H2O inlet partial pressure
8
9
10 of 160 kPa. Most of the data only exhibit a minor dependence on the CH4 conversion. The two empty
11
12
13 symbols deviate more strongly from the simulated lines and are considered as outliers. They were obtained
14
15
16 with a CH4 inlet partial pressure of 20 or 40 kPa which were the lowest pressures used, so produced the
17
18
19 lowest amounts of CO and CO2. Therefore, these points were retained in this figure, but omitted in the
20
21
22 regression analysis.
23
24
25 100
26
27
28
29 80
Selectivity / %

30
31
32 60
33
34
35
36 40
37
38
39 20
40
41
42
43 0
44 30 40 50 60
45 CH4 conversion / %
46
47
48 Fig. 5. Selectivity to CO and CO2 as a function of the conversion for the simple model at 0.40 MPa and
49 923 K, a catalyst mass of 0.1 g, a H2O inlet partial pressure of 160 kPa and a CH4 inlet partial pressure
50
51 between 20 and 100 kPa. Points: experimentally observed, lines: calculated by solving the set of Eq. 8
52 with the net rates of formation given by Eqs. 19 to 23. The rate and equilibrium coefficients are estimated
53
54 by a weighted regression and are shown in Table 2. , full line, black - selectivity to CO, , dashed line,
55 red - selectivity to CO2.
56
57
58
59 26
60
ACS Paragon Plus Environment
Page 27 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
5
Fig. 6 shows a parity diagram for each of the responses for the simple model. The weighted regression
6
7 shows an excellent agreement between model simulated and experimentally measured CH4 outlet molar
8
9
flow rates, a good agreement for H2O, CO, and CO2, and moderate agreement for H2. The closure of the
10
11
12 carbon balance of the experimental data, as explained in the procedures, has led to a slight but consistent
13
14
overestimation of the H2 outlet molar flow rate and corresponding underestimation of the CO2 outlet molar
15
16
17 flow rate in the experimental dataset. The non-zero average value of the error of two out of five model
18
19 responses, i.e., the outlet flow rate of H2O and H2, is attributed to the non-closure of the hydrogen and
20
21
22 oxygen balance in the experimental data.
23
24
25 100 220
26
Predicted FCH / 10-6 mol s-1

Predicted FH O / 10-6 mol s-1

27 80 180
28
29
60 140
30
31
4

32 40 100
2

33
34 20 60
35
36 0 20
37 0 20 40 60 80 100 20 60 100 140 180 220
38 -6 -1 -6 -1
Observed FCH / 10 mol s Observed FH O / 10 mol s
39 4 2

40
41 10 24
42
Predicted FCO / 10-6 mol s-1
-1

22
Predicted FCO / 10 mol s

43 8
44 20
45
-6

6
46 18
47 16
2

4
48
49 14
50 2
12
51
52 0 10
53 0 2 4 6 8 10 10 12 14 16 18 20 22 24
-6 -1 -6 -1
54 Observed FCO / 10 mol s Observed FCO / 10 mol s
2
55
56
57
58
59 27
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 28 of 42

1
2
3
4 140
5

Predicted FH / 10-6 mol s-1


6 120
7
8 100
9
10

2
80
11
12
60
13
14
15 40
40 60 80 100 120 140
16
17
Observed FH / 10-6 mol s-1
2

18
19 Fig. 6. Parity diagram for the simple model of the molar outlet flow rate FCH4, FH2O, FCO, FCO2 and FH2 for
20
21 the weighted regression at a temperature of 923 K and a total pressure of 0.4 MPa. The estimated rate and
22 equilibrium coefficients are shown in Table 2.
23
24
25
26
27
28 Table 3 compares results of the simple model and a model that considers the reversibility of the WGS
29
30
31 reaction. Shown are the parameter estimates and the corresponding 95 % confidence intervals; K1 for
32
33
34 CH4 adsorption, K2 for H2O adsorption, k3 for steam methane reforming and k4 for the water-gas shift
35
36
37 reaction. The K1, K2, and k3 parameters are close in both models, indicating no major deviations in their
38
39
40 interpretation. The k4 values are necessarily different because of the assumption of reversibility, which
41
42
43 k 4+ K S
44 introduces the equilibrium constant for the WGS in the calculations ( K 4 = = ).
45
k 4 K 2
46
47
48 For the simple model all parameters are statistically significant as shown by the narrow confidence
49
50
51 intervals. For the results of this paper, the tabulated Fa value for model adequacy is 1.48. The calculated
52
53
54 Fa value of 2.39 is very close to the tabulated one, indicating that the model is on the edge of being
55
56
57 adequate.
58
59 28
60
ACS Paragon Plus Environment
Page 29 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
The tabulated F value with respect to the global significance is 2.42. The calculated Fs value of the
5
6
7
8
weighted regression of 1.02x104 greatly exceeds this value and hence the performed regression is globally
9
10
11 significant. The maximum binary correlation coefficient amounts to -0.9043 between the equilibrium
12
13
14 coefficient K1 and the rate coefficient k3.
15
16
17 For the simple model with consideration of the reverse WGS reaction the tabulated Fa value for model
18
19
20 adequacy is 1.48. The calculated Fa value of 2.84 is very close to the tabulated one, indicating that the
21
22
23 model is also on the edge of being adequate.
24
25
26 The tabulated F value with respect to the global significance is 2.42. the calculated Fs value of the weighted
27
28
29 regression of 7.81x103 greatly exceeds this value and hence the performed regression is globally
30
31
32 significant. The maximum binary correlation coefficient amounts to -0.93 between the equilibrium
33
34
35 coefficient K2 and the rate coefficient k3.
36
37
38 Overall, both models are adequate. The simple model has narrower parameter confidence intervals while
39
40
41
the simple model with the reverse WGS reaction has a larger Fs value.
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 29
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 30 of 42

1
2
3
4 Table 3. Calculated model parameters at 923 K for the weighted regressions with 95 % confidence interval
5
6
7
8 Simple model with
Simple model Units
9 reversible water-gas shift
10
11
12
K1 30.3 0.3 28.8 4.8 MPa-1
13
14
15 K2 13.5 0.1 12.5 2.5 MPa-1
16
17 k 3' = k 3 zC tot 1.7 0.1 1.63 0.17 mol kgcat-1 s-1
18
19 k 4' = k 4 C tot 74.0 0.5 1.11 0.03 mol MPa-1 kgcat-1 s-1
20
21 Fs 1.02x104 7.81x1043 -
22
23 Fa 2.39 2.84 -
24
25
26
27
28
29
30 Fig. 7 shows the calculated coverages of the catalyst surface obtained with the simple sequence by species
31
32 derived from CH4 (CH4), H2O (H2O), and empty sites (*), as well as the product of CH4 and H2O as a
33
34
35 function of the H2O partial pressure at a CH4 partial pressure amounting to 80 kPa and calculated from K3
36
37 and K4 obtained from the weighted regression, see Table 2, and Eqs. 16-18. With increase of the partial
38
39
40 pressure of H2O, the catalyst surface coverage by CH4 decreased, as the coverage by H2O increased. As
41
42 was discussed in the description of the reaction sequence, the coverages denote species associated with
43
44
45 methane (CH4) and species associated with water (H2O), not necessarily the reactants themselves, as these
46
47 would probably not be on the surface at high coverages. Consideration of actual species would greatly
48
49
50 complicate the analysis, introducing many additional parameters, that is unwarranted from the data. The
51
52 product of both coverages reaches a maximum with increasing H2O inlet partial pressure. This reflects the
53
54
55 behavior of the CH4 conversion versus the H2O inlet partial pressure, see Fig. 3. Indeed, Eq. 14 shows that
56
57 the CH4 consumption rate in SMR (r3) is proportional to the product of the coverage of CH4 and H2O.
58
59 30
60
ACS Paragon Plus Environment
Page 31 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
5
Therefore, the model follows this trend, Fig. 3. It can be noted that the fraction of free sites on the surface
6
7 decreases as the H2O partial pressure increases, but falls slightly less than the coverage in CH4 derived
8
9
products.
10
11
12
13
14
15
16 1.0 0.20
17
18 CH x H O
4 2
19 0.8
20 0.15

CH x H O / -
Coverage / -

21
H O

2
22 0.6
23 2

24 0.10
*

4
25 0.4
26
27
28 0.05
0.2
29
30
CH
4

31
0.0 0.00
32
100 200 300
33
34 H2O partial pressure / kPa
35
36
37 Fig. 7. Coverage of catalyst surface by species related to H2O (H2O), CH4 (CH4), empty sites (*) and the
38 product of H2O and CH4 as a function of the partial pressure of H2O. Partial pressure of CH4: 80 kPa,
39
40 Temperature: 923 K, Total pressure: 0.40 MPa. The coverages are calculated with Eqs. 14 to 16. The
41 equilibrium coefficients are estimated by a weighted regression and are shown in Table 2. Full line, black
42
43 - H2O x CH4, dashed line, red - H2O, dotted line, green - *, dash-dot line, blue - CH4.
44
45
46
47
48 The kinetic model with the model parameters estimated by the weighted regression is subsequently used
49
50
51 to simulate the CH4 conversion at a WHSV of 48000 Ncm3 h-1 gcat-1, a H2O/CH4 ratio of 2 and a total
52
53
54 pressure of 0.1 MPa, see filled square in Fig. 2. By decreasing the total pressure from 0.4 MPa to 0.1
55
56
57 MPa, the conversion drops as indicated by the arrow in Fig. 2. This is an interesting result, as the
58
59 31
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 32 of 42

1
2
3
4
equilibrium conversion decreases with increasing pressure due to the stoichiometry of the SMR reaction,
5
6
7
8
see Eq. 1. The increase in conversion with pressure is due to enhanced kinetics, with rates increasing
9
10
11 with higher coverages of reactants as given by Eqs. 11 and 12. This will be discussed subsequently.
12
13
14
15 An important result from the model developed in this work is presented in Fig. 8 which illustrates the
16
17
18 effects of total pressure and space time on the CH4 conversion. The currently developed model did
19
20
21 not include thermodynamic constraints as it assumed an irreversible reaction, hence, an additional
22
23
24 software package (NASA, Chemical Equilibrium with Applications) was used to determine the
25
26
27 equilibrium conversion. As expected, the CH4 conversion decreases with increasing inlet CH4 flow rate.
28
29
30 At low space velocity the conversion is lower at higher pressure because of thermodynamic equilibrium.
31
32
33 This is, of course, because the increase in moles in the SMR dictates that conversion will be lower at
34
35
36 higher pressure by Le Chateliers principle. However, at higher space velocity a transition is made to
37
38
39 the kinetic regime. Thus, a higher total pressure leads to a higher reaction rate, because adsorption rates
40
41
42
43
are higher, as long as the reaction is not close to thermodynamic equilibrium. These results are not
44
45
46 due to the suppression of the back reaction, as the model used here assumes an irreversible reaction
47
48
49 step. Those results rationalize why SMR is practiced industrially at high pressure, despite the
50
51
52 unfavorable equilibrium. The decrease in conversion with space velocity is reported in other
53
54
55 studies24,25,36,37, but the crossover has not been noted.
56
57
58
59 32
60
ACS Paragon Plus Environment
Page 33 of 42 Industrial & Engineering Chemistry Research

1
2
3
4 Space time / kgcat s molCH -1
5 4

6 -- 8.07 4.03 2.69 2.02 1.61


7 100
8 0.10 MPa (Xeq= 86.7%)
9
10 80 0.20 MPa(Xeq= 74.3%)
CH4 conversion / -

11
12
13 0.40 MPa(Xeq= 60.6%)
14 60
15
16
17 40
18
19
20
21 20
22
23
24
Thermodynamic control Kinetic control
25
0
26 0 10000 20000 30000 40000 50000
3 -1 -1
27 WHSV / Ncm CH4
g h
cat
28
29
30
31 Fig. 8. Simulation result of CH4 conversion change as a function of the WHSV/space time of CH4 at
32
various pressures. The points are the result of the calculation with Eq. 19 and Eq. 20 and the lines are
33
34 drawn by B-spline through the points including the equilibrium points, temperature: 923 K, inlet gas
35
36
composition: H2O/CH4/N2 = 2/1/2, total inlet flow of 400 Nml min-1. The dashed lines show the
37 equilibrium CH4 conversion for each reaction condition. The rate and equilibrium coefficients are
38
39 estimated by a weighted regression and are shown in Table 2. Full line, black 0.1 MPa, dash-dot line -
40 red, 0.2 MPa, dashed line, blue 0.4 MPa.
41
42
43
44
Fig. 9 shows the effect of changing the H2O/CH4 ratio on the CH4 conversion as a function of WHSV.
45
46
47 The CH4 conversion is higher for higher H2O/CH4 ratios and decreases with increasing WHSV. These
48
49
50 trends are similar to those seen in Fig. 2 for both higher and lower H2O/CH4 ratio and can be understood
51
52
53 based on the analysis of the surface coverages as shown in Fig. 7. Thus, the model developed in this
54
55
56 work is able to describe general trends found from analysis of literature data.
57
58
59 33
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 34 of 42

1
2
3
4 Space time / kgcat s molCH -1
5 16.14 13.45 11.53 4
10.09
6 100
7
8 3.0
9 80
10 CH4 conversion / - 2.0
11
12 1.5
60
13
14 1.0
15 40
16 H2O/CH4 = 0.5
17
18
19
20
20
21
22 0
23 5000 6000 7000 8000
24
WHSV / Ncm3CH h-1 gcat-1
4

25
26
27 Fig. 9. Comparison of simulated CH4 conversions as a function of WHSV/space time of CH4 varying
28 H2O/CH4 ratios at 923 K, 0.10 MPa, total inlet flow of 400 Nml min-1, inlet flow CH4 80 Nml min-1. The
29
30 lines are drawn by B-spline through each points. The rate and equilibrium coefficients are estimated by a
31
weighted regression and are shown in Table 2. Full line, magenta - H2O/CH4 = 3.0, green - H2O/CH4 =
32
33 2.0, blue- H2O/CH4 = 1.5, red - H2O/CH4 = 1.0, black - H2O/CH4 = 0.5.
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 34
60
ACS Paragon Plus Environment
Page 35 of 42 Industrial & Engineering Chemistry Research

1
2
3
4 4. CONCLUSIONS
5
6
7 A Ni/MgO-SiO2 catalyst was used to study the catalytic steam methane reforming in a tubular reactor
8
9
10 at 0.40 MPa and 923 K. The CH4 and H2O partial pressure effects were measured and analyzed by a
11
12
13 combined simple reaction model that considered the adsorption of the reactants and the surface reaction
14
15
16 of their derived species, as well as the water-gas shift reaction . Values for the rate and equilibrium
17
18
19 coefficients in this model were obtained by regression and resulted in a good agreement with
20
21
22
experimental results as well as generally reported trends in the literature. Consideration of the
23
24
25
26
reversibility of the water-gas shift reaction did not significantly improve the simple model.
27
28
29 Particularly interesting was that the space velocity (space time) effect on the methane conversion
30
31
32 revealed a cross-over with total pressure due to transition from thermodynamic to kinetic control. Due
33
34
35 to the molar expansion during reaction, increasing the total pressure will lead to a decrease in conversion
36
37
38 when operating in the thermodynamic controlled regime. In the kinetic controlled regime, the reactant
39
40
41 surface coverages monotonically increase with total pressure, resulting in higher methane conversions.
42
43
44 Model predictions at lower total pressures correspond to experimental observations reported in the
45
46
47 literature, confirming that the developed model, although extremely simple, conveys the most essential
48
49
50 features for SMR.
51
52
53
54
55
56
57
58
59 35
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 36 of 42

1
2
3
4 NOMENCLATURE
5
6
7 Roman
8 A matrix A -
9
10 b estimated model parameter -
11
CA concentration of A mol m-3
12
13 concentration of A adsorbed on
14 CA* mol kg-1
15
surface
16 total concentration of active
17 Ctot mol kg-1
18 sites
19 C dimensionless concentration -
20
21 dp catalyst pellet diameter m
22 De effective gas-phase diffusivity m2 s-1
23
24 e residual mol s-1
25 Fa F value for adequacy -
26
27 Fs F value for significance -
28 FA molar flow rate of A mol s-1
29
30 H0 standard reaction enthalpy kJ mol-1
31
32
i.d. inner diameter m
33 k reaction rate coefficient reaction dependent
34
35 kc mass transfer coefficient m s-1
36 K equilibrium coefficient Pa-1
37
38 n reaction order -
39 o.d. outer diameter m
40
41 p pressure Pa
42 r specific reaction rate mol kgcat-1 s-1
43
44 RA specific production rate of A mol kgcat-1 s-1
45 R universal gas constant J mol-1 K-1
46
47 T temperature K
48
W catalyst mass kg
49
50 X independent variable -
51
52
X conversion %
53 y measured quantity mol s-1
54
55 z number of neighbors -
56 Greek
57
58
59 36
60
ACS Paragon Plus Environment
Page 37 of 42 Industrial & Engineering Chemistry Research

1
2
3
4 model parameter -

5 experimental error mol s-1


6
7 A surface coverage of A -
8 density kg m-3
9
10 Subscripts
11
12
b bulk -
13 cat catalyst -
14
15 eq equilibrium -
16 Superscripts
17
18 0 standard, inlet -
19 * active site -
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 37
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 38 of 42

1
2
3
4 AUTHOR INFORMATION
5
6 Corresponding authors
7
8
9
10 S. Ted Oyama Email oyama@vt.edu
11
12 Joris Thybaut Email joris.thybaut@ugent.be
13
14 Notes
15
16
17
18 The authors declare no conflicting interests.
19
20
21 ACKNOWLEDGEMENTS:
22
23
24
25
26
This work was supported by a doctoral fellowship from the Fund for Scientific Research Flanders
27
28
29 (FWO), the European Research Council under the European Unions Seventh Framework Programme
30
31
32 (FP7/2014-2019) / ERC grant agreement n 615456 and the Long Term Structural Methusalem
33
34
35 Funding by the Flemish Government. NK is grateful for support from Kagaku Jinzai Ikusei Program.
36
37
38 STO received support from the director, National Science Foundation under grant CHE-1361842. AT
39
40
41 and STO also received assistance from Fuzhou Univ., College of Chemical Engineering. The catalyst
42
43
44 was kindly supplied by JGC Catalysts and Chemicals Ltd.
45
46
47
48
49 SUPPORTING INFORMATION
50
51 Brief description of the simple model fit augmented by reversibility of the water-gas shift reaction.
52
53
54 Comparison of results of the simple model and the model with reversibility of the WGS reaction.
55
56
57 Details of the calculation of the Mears Criterion for lack of heat transfer limitations.
58
59 38
60
ACS Paragon Plus Environment
Page 39 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
5
6
7
8
9
10 REFERENCES
11
12
( 1 ) Chatrattanawet, N.; Skogestad, S.; Arpornwichanop, A. Control structure design and dynamic
13
14 modeling fora solid oxide fuel cell with direct internal reforming of methane. Chem. Eng. Res. Des. 2015,
15
16
98, 202.
17 (2) Kang, K.-S.; Kim, C.-H.; Bae, K.-K.; Cho, W.-C.; Jeong, S.-U.; Lee, Y.-J.; Park, C.-S. Reduction and
18
19 oxidation properties of Fe2O3/ZrO2 oxygen carrier for hydrogen production. Chem. Eng. Res. Des. 2014,
20 92, 2584.
21
22
(3) Hacarlioglu, P.; Gu, Y.; Oyama, S.T., Dry reforming of methane has no future for hydrogen
23 production: Comparison with steam reforming at high pressure in standard and membrane reactors.
24
25 Int. J. Hydr. Energy 2012, 37, 10444-10450.
26
27 (4) Flaherty, D. W.; Yu, W.-Y.; Pozun, Z. D.; Henkelman, G.; Mullins, C. B. Mechanism for the watergas
28
29 shift reaction on monofunctional platinum and cause of catalyst deactivation. J. Catal. 2011, 282, 278.
30 (5) Zheng, R.; Diver, R.; Caldwell, D.; Fritz, B.; Cameron, R.; Humble, P.; TeGrotenhuis, W.; Dagle, R.;
31
32 Wegeng, R. Integrated solar thermochemical reaction system for steam methane reforming. Energy
33 Procedia 2015, 69, 1192.
34
35 (6) Hafizi, A.; Rahimpour, M. R.; Hassanajili, S. Calcium promoted Fe/Al2O3 oxygen carrier for hydrogen
36 production via cyclic chemical looping steam methane reforming process. Int. J. Hydrogen Energy 2015,
37
38 40, 16159.
39 (7) Vaiano, V.; Sacco, O.; Pisano, D.; Sannino, D.; Ciambelli, P. From the design to the development of a
40
41 continuous fixed bed photoreactor for photocatalytic degradation of organic pollutants in waste water.
42
Chem. Eng. Sci. 2015, 137, 152.
43
44 (8) Sadooghi, P.; Rauch, R. Experimental and modeling study of hydrogen production from catalytic steam
45
46
reforming of methane mixture with hydrogen sulfide. Int. J. Hydrogen Energy 2015, 40, 10418.
47 (9) Chen D; Lodeng R; Svendsen H; Holmen A. Hierarchical multiscale modeling of methane steam
48 reforming reactions. Ind Eng Chem Res. 2011, 50,2600-12.
49 (10) Foppa L; Silaghi MC; Larmier K; Comas-Vives A. Intrinsic reactivity of Ni, Pd and Pt surfaces in
50 dry reforming and competitive reactions: Insights from first principles calculations and microkinetic
51 modeling simulations. J Catal. 2016, 343, 196-207.
52 (11) Sprung C; Kechagiopoulos P,N; Thybaut J,W; Arstad B; Olsbye U; Marin G,B. Microkinetic
53 evaluation of normal and inverse kinetic isotope effects during methane steam reforming to synthesis
54 gas over a Ni/NiAl2O4 model catalyst. Applied Catalysis A: Gen. 2015, 492, 231-42.
55
56
(12) Chung, S.-H.; Angelici, C.; Hinterding, S.O.M.; Weingarth, M.; Baldus, M.; Houben, K.;
57 Weckhuysen, B.M.; Bruijnincx, P.C.A. Role of magnesium silicates in wet-kneaded silicamagnesia
58
59 39
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 40 of 42

1
2
3
4
5
catalysts for the Lebedev ethanol-to-butadiene process, ACS Catal. 2016, 6, 4034-4045.
6 (13) Kitayama, Y.; Satoh, M.; Kodama, T. Preparation of large surface area nickel magnesium silicate
7 and its catalytic activity for conversion of ethanol into buta-1,3-diene, Catal. Lett. 1996, 36, 95-97.
8 (14) Oyama, S. T.; Zhang, X.; Lu, J.; Gu, Y.; Fujitani, T. Epoxidation of propylene with H2 and O2 in the
9
10 explosive regime in a packed-bed catalytic membrane reactor. J. Catal. 2008, 257, 1.
11 (15) Toch, K.; Thybaut, J. W.; Marin, G. B. A systematic methodology for kinetic modeling of chemical
12
13 reactions applied to n-hexane hydroisomerization. AIChE J. 2015, 61, 880.
14
(16) Constales, D.; Yablonsky, G.S.; D'hooge, D.; Thybaut, J.W.; Marin, G.B. Advanced Data Analysis
15
16 and Modelling in Chemical Engineering, First ed.; Elsevier: Amsterdam, 2016.
17
(17) Stewart, W.E.; Caracotsios, M. Computer-Aided Modeling of Reactive Systems; John Wiley & Sons,
18
19 Inc.: New York, 2008.
20
21 (18) Metaxas, K.; Thybaut, J. W.; Morra, G.; Farrusseng, D.; Mirodatos, C.; Marin, G. B. A microkinetic
22 vision on high-throughput catalyst formulation and optimization: development of an appropriate software
23
24 tool. Top. Catal. 2010, 53, 64-76.
25 (19) Yaakob, Z.; Bshish, A.; Ebshish A., Tasirin, S.M.; Alhasan, F.H. Hydrogen production by steam
26
27 reforming of ethanol over nickel catalysts supported on sol gel made alumina: Influence of calcination
28
29 temperature on supports, Materials 2013, 6, 2229
30
31 (20) Wu, H.; La Parola, V.; Panteleo, G.; Puleo, F.; Venezia, A. M.; Liotta, L.F. Ni-based catalysts for low
32 temperature methane steam reforming: recent results on Ni-Au and comparison with other bi-metallic
33
34 systems. Catalysts 2013, 3, 563.
35 (21 ) Karimipourfard, D.; Kabiri, S.; Rahimpour, M. R. A novel integrated thermally double coupled
36
37 configuration for methane steam reforming, methane oxidation and dehydrogenation of propane. J. Nat.
38 Gas Sci. Eng. 2014, 21, 134.
39
40 (22 ) Xu, P.; Zhou, Z.; Zhao, C.; Cheng Z. Catalytic performance of Ni/CaO-Ca5Al6O14 bifunctional
41 catalyst extrudate in sorption-enhanced steam methane reforming. Catal. Today 2016, 259, 347.
42
43 (23) Kim, T. W.; Park, J. C.; Lim, T.-H.; Jung, H.; Chun, D. H.; Lee, H. T.; Hong, S.; Yang, J.-I. The
44
kinetics of steam methane reforming over a Ni/-Al2O3 catalyst for the development of small stationary
45
46 reformers. Int. J. Hydrogen Energy 2015, 40, 4512.
47
48
(24) Lee, S. Y.; Lim, H.; Woo, H. C. Catalytic activity and characterizations of Ni/K2TixOy-Al2O3 catalyst
49 for steam methane reforming. Int. J. Hydrogen Energy, 2014, 39, 17645.
50
51 (25) Chu, B.; Zhang, N.; Zhai, X.; Chen, X.; Cheng, Y. Improved catalytic performance of Ni catalysts for
52 steam methane reforming in a micro-channel reactor. J. Energy Chem. 2014, 23, 593.
53
54 (26) Boukha, Z.; Jimnez-Gonzlez, C.; de Rivas, B.; Gonzlez-Valasco, J. R.; Gutirrez-Ortiz, J. I.;
55 Lpez-Fonseca, R. Synthesis, characterisation and performance evaluation of spinel-derived Ni/Al2O3
56
57 catalysts for various methane reforming reactions. Appl. Catal., B 2014, 158-159, 190.
58 (27) Kho, E. T.; Scott, J.; Amal, R. Ni/TiO2 for low 40
temperature steam reforming of methane. Chem. Eng.
59
60
ACS Paragon Plus Environment
Page 41 of 42 Industrial & Engineering Chemistry Research

1
2
3
4
5 Sci. 2016, 140, 161.
6 (28) Zhang, Y.; Wang, W.; Wang, Z.; Zhou, X.; Wang, Z.; Liu, C.-J. Steam reforming of methane over
7
8 Ni/SiO2 catalyst with enhanced coke resistance at low steam to methane ratio. Catal. Today 2015, 256,
9 130.
10
11 (29) Roh, H.-S.; Eum, I.-H.; Jeong, D.-W. Low temperature steam reforming of methane over Ni-Ce(1-
12
x)Zr(x)O2
catalysts under severe conditions. Renew. Energy 2012, 42, 212.
13
14 (30) Wu, X.; Wu, C.; Wu, S. Dual-enhanced steam methane reforming by membrane separation of H2 and
15
16
reactive sorption of CO2. Chem. Eng. Res. Des. 2015, 96, 150.
17 (31) Gil, A. G.; Wu, Z.; Chadwick, D.; Li, K. Ni/SBA-15 catalysts for combined steam methane reforming
18
19 and water gas shiftPrepared for use in catalytic membrane reactors. Appl. Catal., A 2015, 506, 188.
20 (32) Hungria, A. B.; Fernndez-Garcia, M.; Anderson, J. A.; Martnez-Arias, A. The effect of Ni in Pd
21
22 Ni/(Ce,Zr)Ox/Al2O3 catalysts used for stoichiometric CO and NO elimination. Part 2: Catalytic activity
23 and in situ spectroscopic studies. J. Catal. 2005, 235, 262.
24
25 33) Marin, G. B.; Yablonsky, G. S. Kinetics of Chemical Reactions; John Wiley & Sons, Inc.: New York,
26 2011.
27
28 (34) G. Djega-Mariadassou, M. Boudart, Kinetics of Heterogeneous Catalytic Reactions; Princeton
29
University Press: Princeton, 1984.
30
31 (35) M.A. Vannice, Kinetics of Catalytic Reactions, Springer, New York, 2005.
32 (36) Kyriakou, V.; Garagounis, I.; Vourros, A.; Vasileiou, E.; Manerbino, A.; Coors, W.G.; Stoukides,
33
34 M., Methane steam reforming at low temperatures in a BaZr0.7Ce0.2Y0.1O2.9 proton conducting
35
36 membrane reactor. Appl. Catal. B: 2016, 186, 1.
37
38 (37) Boeltken, T.; Wunsch, A.; Gietzelt, T.; Pfeifer, P.; Dittmeyer, R., Ultra-compact microstructured
39 methane steam reformer with integrated palladium membrane for on-site production of pure hydrogen:
40
41 Experimental demonstration. Int. J. Hydrogen. Energy 2014, 39, 18058.
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59 41
60
ACS Paragon Plus Environment
Industrial & Engineering Chemistry Research Page 42 of 42

1
2
3
4
5 Crossover of CH4 conversion curves with WHSV/space time due to a transition between
6
7 thermodynamic to kinetic control.
8
9
10
11 Space time / kgcat s molCH -1
4
12 -- 8.07 4.03 2.69 2.02 1.61
13 100
14
15 0.10 MPa (Xeq= 86.7%)
16
80 0.20 MPa(Xeq= 74.3%)
CH4 conversion / -

17
18
19 0.40 MPa(Xeq= 60.6%)
20 60
21
22
23
24 40
25
26
27 20
28
29
30 Thermodynamic control Kinetic control
31 0
32 0 10000 20000 30000 40000 50000
33 3 -1 -1
34
WHSV / Ncm CH4
g h
cat
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
ACS Paragon Plus Environment

You might also like