You are on page 1of 12

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 1 7 9 8 e1 1 8 0 9

Available at www.sciencedirect.com

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

Reaction/separation coupled equilibrium modeling of steam


methane reforming in fluidized bed membrane reactors
Donglai Xie*, Weiyan Qiao, Ziliang Wang, Weixing Wang, Hao Yu, Feng Peng
MOE Key Laboratory of Enhanced Heat Transfer & Energy Conservation, South China University of Technology, Guangzhou 510640, China

article info

abstract

Article history:

An equilibrium model of steam methane reforming coupled with in-situ membrane sepa-

Received 29 May 2010

ration for hydrogen production was developed. The model employed Sieverts Law for

Received in revised form

membrane separation and minimum Gibbs energy model for reactions. The reforming and

25 August 2010

separation processes were coupled by the mass balance. The model assumed a continu-

Accepted 28 August 2010

ously stirred tank reactor for the fluidized bed hydrodynamics. The model predictions for

Available online 20 September 2010

a typical case were compared with those from the model of Ye et al. [15] which assumed
a plug flow for bed hydrodynamics. The model predictions show satisfactory agreement

Keywords:

with experimental data in the literatures. The influences of reactor pressure, temperature,

Hydrogen

steam to carbon ratio, and permeate side hydrogen partial pressure on solid carbon, NHx

Fluidized bed

and NOx formation were studied using the model.

Reforming

2010 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved.

Membrane
Modeling

1.

Introduction

An energy economy based on hydrogen could alleviate the


worlds growing concerns about energy supply, security, air
pollution, and greenhouse gas emissions. Hydrogen offers
long-term potential for an energy system that produces nearzero emissions based on each countrys domestically available resources. It is also a major industrial commodity, used as
an intermediate in a number of chemical and metallurgical
processes, for example in the production of ammonia and
methanol, upgrading of heavy hydrocarbon feed stocks, iron
ore reduction and food processing [1].
Most of the worlds hydrogen supply is generated by steam
reforming or partial oxidation of natural gas in parallel fixed bed
reactors within huge top-fired or side-fired furnaces, coupled
with Pressure Swing Adsorption (PSA) for hydrogen purification
[2]. Although this technology has been widely used for decades,
it still suffers from several disadvantages such as low catalyst

effectiveness, low heat transfer rates, large temperature gradients within the bed and thermodynamic equilibrium constraint
on chemical reaction [3e6]. To overcome these problems,
Fluidized Bed Membrane Reactors (FBMR) have been proposed
and developed by several research groups [1,4e7].
The principal reactions involved in catalytic steam
methane reforming are [8]:Steam methane reforming (SMR):
CH4 H2O CO 3H2,

DHo298 206 kJ/mol

(R1)

CH4 2H2O CO2 4H2,

DHo298 165 kJ/mol

(R2)

Wateregas shift (WGS):


CO H2O CO2 H2,

DHo298 41 kJ/mol

(R3)

The first two reactions are strongly endothermic, and both


lead to significant increase in molar flow rates as the reaction

* Corresponding author. Tel./fax: 86 20 22236985.


E-mail address: dlxie@scut.edu.cn (D. Xie).
0360-3199/$ e see front matter 2010 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2010.08.130

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 1 7 9 8 e1 1 8 0 9

proceeds. Equilibrium conversions of both reforming reactions benefit from high temperatures and low pressures,
whereas the wateregas shift reaction (R3), being exothermic
and having no change in the number of molars, benefits
thermodynamically from lower temperatures and is independent of pressure.
To sustain the above endothermic reactions (R1) and (R2),
oxygen or air can be introduced into the system with the
following oxidation reactions taking place [9]:

11799

C H2O CO H2,

DHo298 131 kJ/mol

C 0.5O2 CO,

DHo298 111 kJ/mol

(R10)

production process in fluidized bed membrane reactors in


literature. The model of Grace et al. [14] considered a Continuously Stirred Tank Reactor (CSTR) for the bed hydrodynamics, and pure oxygen for autothermal operation. Solid
carbon formation was studied using the model. The amount of
hydrogen separated by membrane was used as an input for
the model. Hence the reaction process and separation process
was not coupled. The model of Ye et al. [15] considered the
FBMR as a Plug Flow Reactor (PFR) for reactor hydrodynamics.
It did not consider the oxidant addition to the bed, nor solid
carbon formation as product. The coupling of the chemical
reaction and membrane separation processes was achieved
by a sequential modular approach. The FBMR was divided into
several successive steam methane sub-reformers and
membrane sub-separators. The process is represented by
(m 1) sub-reformers and m sub-separators. The reactor offgases from the ith sub-reformer are fed to the ith sub-separator, the non-permeated gases from the ith sub-separator are
fed to the (i 1)th sub-reformer, and the permeated hydrogen
from the ith sub-separator accumulates in the (i 1)th subseparator. Hence the coupling of the reaction and separation
process is a pseudo-coupling.
In practice, if autothermal operation is preferred, it is
better to use air, rather than pure oxygen as oxidant. When air
is introduced, the following side reactions involving nitrogen
could take place:

C CO2 2CO,

DHo298 172 kJ/mol

(R11)

N2 3H2 2NH3

DHo298 98 kJ/mol

(R12)

Modeling of FBMR presents interesting challenges because


of the coupling of selective diffusion through the permeable
membrane with chemical reactions and mass transfer on the
reactor side [12]. It is useful to investigate the effects of key
operating parameters on the reactor performance, and also
the performance of the reactor system can be explored beyond
the range of parameters that can be studied experimentally
due to limitations imposed by economic and safety
considerations.
Two types of methods can be tried for the modeling of
chemical reactions, the kinetic approach and thermodynamic
approach. Many kinetic models have been proposed to simulate the steam methane reforming process for hydrogen
production, whose features have been summarized well by
Xie [13]. Predictions from these models were claimed by their
authors to be in good agreement with experimental data.
However, Kinetic models are limited to small numbers of
reactions and species with clearly defined mechanism. For
complex systems, the reaction mechanisms often require
extensive study. Also these models were solved by FORTRAN,
Matlab or other computer programs, which are not easily
accessible to design engineers in industry. For operating
conditions of interest (i.e., temperature 550e900  C and
absolute pressure 0.5e3.0 MPa), the steam reforming and
wateregas shift reactions, or the oxidation reactions in the
presence of catalyst are fast enough that the production of
hydrogen closely approaches the equilibrium values [14]. As
a result, thermodynamic equilibrium analysis provides
a simple and direct basis for practical applications. To the
knowledge of the authors, two thermodynamic equilibrium
models have been developed to simulate the hydrogen

N2 2H2 N2H4

DHo298 95 kJ/mol

(R13)

N2 2O2 2NO2

DHo298 68 kJ/mol

(R14)

N2 O2 2NO

DHo298 181 kJ/mol

(R15)

N2 2O2 N2O4

DHo298 10 kJ/mol

(R16)

N2O2 2N2O

DHo298 163 kJ/mol

(R17)

CH4 0.5O2 CO 2H2,

DHo298 36 kJ/mol

(R4)

CH4 1.5O2 CO 2H2O,

DHo298 607 kJ/mol

(R5)

CH4 2O2 CO2 2H2O,

DHo298 802 kJ/mol

(R6)

Side reactions, like the catalytic dry reforming of methane,


methane cracking, etc. could also present accompanying the
above principle reactions, including [9e11]:
CH4 CO2 2CO 2H2,
CH4 C 2H2,

DHo298 247 kJ/mol


DHo298 75 kJ/mol

(R7)
(R8)
(R9)

The possibility of the formation of NOx and NHx with the


presence of nitrogen needs to be studied, which has not been
considered previously.
In this work, a thermodynamic model is developed to
model the fluidized bed membrane reactor for hydrogen
production. In the model, the hydrogen separation process by
the membrane is coupled with the steam methane reforming
process with or without air addition for heat supply. The
formation of solid carbon, NOx and NHx was also considered.

2.

Model development

2.1.

Primary assumptions

A fluidized bed membrane reactor for pure hydrogen production by steam methane reforming is shown schematically in
Fig. 1. Preheated high-temperature (usually 500e800  C) and
high-pressure (usually 1e3 MPa) natural gas and steam are
premixed and fed into the reactor. The reactor contains Nickel
based catalyst and palladium (or its alloy) membrane modules

11800

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 1 7 9 8 e1 1 8 0 9

in tubular or planner shapes. Usually palladiumesilver or


palladiumecopper membrane modules are widely employed
due to their relatively high permeability and longevity
comparing to pure palladium modules. For these membrane
modules the reactor temperature should be controlled below
700  C to avoid damaging the membranes. For autothermal
operations, air is introduced into the reactor to produce the
heat for sustaining the endothermic reforming reactions.
Otherwise, external heat needs to be supplied to the reactor. To
reduce the permeate side hydrogen partial pressure, sweep gas
(usually steam) can be introduced into the permeate side of the
membrane modules. To simplify the simulation of steam
methane reforming and hydrogen separation processes in the
FBMR, the domain sketched by the dashed box in Fig. 1 is
considered for model development.
To represent the characteristics of fluidized bed reactors,
the model assumes:
1. The system is at steady state and reaches thermodynamic
equilibrium.
2. The reacting gas and catalyst are perfectly mixed, i.e., the
reactor is a continuously stirred tank reactor.
3. Uniform temperature within the fluidized bed.
4. Pressure gradients are ignored both within the reactor bed
and within the membrane.
5. Heat losses are negligible.
6. Ideal gas laws are applicable to all reaction gases.
7. To make the model more general, consider CH4, H2O, O2 and
N2 in the feed stream. The reactor of gas constitutes of 15
species, including CO, CO2, H2O, CH4, H2, O2, N2, C2H2, C2H4,
NO, NO2, N2O, NH3, N2H4 and solid carbon.
8. Only hydrogen can penetrate through the membrane i.e.,
only hydrogen and sweep gas are considered in the
permeate side.



EP

 M

RT
QH h k Cep PRH  PM
MH e

Hydrogen + Sweep gas

(1)

The hydrogen flux follows the Sieverts law when the


hydrogen pressure exponent M is equal to 0.5, which is usually
valid for thick Pd films [17]. Deviations from the Sieverts law
(M > 0.5) were reported for very thin membranes [18,19]. Based
on a hydrogen permeation model, Ward and Dao [20] showed
that at temperatures above 400  C, M was equal to 0.5 for
membranes thicker than 1 mm. Usually to use Sieverts Law
correctly with exponent 0.5, the thickness of membrane
should be higher than 10 mm [17]. The co-existence of H2O, CO,
CO2 or CH4 has been reported to have a negative influence on
the membrane separation performance [21]. This negative
effect is considered in the permeation efficiency factor e h,
together with some other factors influencing the membrane
performance, like the existence of a membrane substrate. In
practice, h should be determined experimentally. h is reported
in literature to be from 0.39 to nearly 1.0 [3,12,22].
10. Although solid carbon is considered in the current model,
the purpose is to study the operating regime that the solid
carbon may appear and which should be avoided in practical applications. Simulation results in the following part
of this study show that solid carbon is formed under some
extreme conditions. For fluidized beds, carbon formation
is not concerned under regular operation conditions due to
the high bed temperature uniformity [15]. Hence the
influence of carbon accumulation on membrane and
catalysts on their performances is not considered in the
current model.

2.2.

Non-permeate product gas

Domain considered in the


model

9. Hydrogen permeation through the membrane follows Sieverts Law [16], i.e.:

Governing equations

The most commonly used function to identify the equilibrium


state is Gibbs free energy, which is a suitable parameter to
calculate the equilibrium compositions of the reaction
system. The equilibrium composition brings in temperature
dependence, without requiring detailed information
regarding the specific reaction or catalyst performance.
As the system reaches thermodynamic equilibrium, the
total Gibbs free energy of the system should reach minimum
under the operation temperature. Hence

Membranes module

vGt
0;
vni

Sweep gas

Methane & steam

Air

Fig. 1 e Schematic of a fluidized bed membrane reactor.

i 1e15

(2)

i 1e15 represents the fifteen species in the system, in the


order of CO, CO2, H2O, CH4, H2, O2, N2, C2H2, C2H4, NO, NO2,
N2O, NH3, N2H4 and solid carbon.
The elements of C, O, N, and H conserve prior to and after
the reactions. Hence for the elements of C, O and N,
15
X
i1

ni aik Ak ;

k 1; 2; 3

(3)

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 1 7 9 8 e1 1 8 0 9

k 1, 2, 3 represents the three elements of C, O and N,


respectively.
For element of H, consider part of the hydrogen in the
reacting gas permeates through the membrane:
15
X

ni ai;H  2QH AH

(4)

i1

The total Gibbs free energy of the system is linked to the


composition of the system by
Gt

14
X

ni Gi nC Gc

(5)

i1

Gi is related to the system pressure and temperature by


yi P
Gi G0i RT ln
P0

(6)

where
yi

ni
14
X

(7)
ni

i1

For solid carbon, its Gibbs free energy is related to the system
pressure by
GC VC P  P0

(8)

Where VC is the molar volume of solid carbon and equals to


4.58  106 m3/mol [23].

2.3.

Solution of the governing equations

Using Gibbs free energy minimization approach to predict the


product composition of a reaction system that reaches thermodynamic equilibrium is a well-established method. The
mathematical aspects of the method are well documented
[24e28]. It has been widely used in recent years for process
optimization in energy sectors, for example, gasification
[29e32] and hydrogen production from reforming related
processes [32e37]. In the current study, the model defined by
control equations of (1)e(8) was solved with two approaches.
A) Solved in Matlab with NewtoneRaphson algorithm. With
some modifications on the mass balance of hydrogen element
considering its permeation through membranes, the computational program developed in the current study is following
the work of Li et al. [30], where detailed algorithm information
is described. B) Solved in Microsoft Excel using its solver tool.
In Excel, the minimum Gibbs energy problem becomes an
optimization problem. In the programming solver tool, the
Gibbs energy of the reactor product as a function of the system
pressure, temperature and molar flow rate of product is set as
an objective to reach minimum value. The variables are the
molar flow rates of product gases and solid carbon. The
constraints are those mass balance equations (3) and (4).
Thermodynamic properties of the species involved were
tabled in a separate work sheet and can be accessed by the
solver tool. By employing the solver tool, Excel can automatically adjust the product molar flow rates to make the Gibbs
energy of the system reaching minimum. In both approaches,
the thermodynamic data of G0i for gases were obtained from

11801

JANAF Thermo Chemical Tables [38]. Both approaches can


give identical result.

2.4.

Comparisons with PFR model of Ye et al. [15]

Strong gas back mixing exists in bubbling fluidized beds.


However, gases are still flowing upwards in the bed. Either PFR
or CSTR hydrodynamic models, or combination of these two
models, has been employed in the reformer models. If no
membrane modules were employed for in-situ hydrogen
separation, the thermodynamic models with either PFR or
CSTR hydrodynamic sub-models should give identical
reformer performance predictions. When membrane modules
were employed, the driving force for hydrogen permeation
through membrane is related to the reactor hydrodynamic.
Hence the reactor performance depends on the selection of
hydrodynamic sub-models.
Consider a fluidized membrane reactor that has feed gas
streams of CH4 and steam. In-situ hydrogen separation
modules with membrane thickness of 10 mm and effective
surface area varying from 0 to 1.8 m2 are installed inside the
reactor. The corresponding Cep varies from 0 to 180 Km. The
base operation condition is P 2.0 MPa, T 650  C,
FCH4 226 mol/h, FS 0, SC 3.0, OC 0 mol/h, P 2.0 MPa,
PM 0.1 MPa. The model of Ye et al. [15] was used to predict the
reactor performance under the conditions of plug flow, and it
was compared with the predictions from the current model for
the conditions assuming CSTR for reactor hydrodynamics. The
reactor performance was denoted by two parameters: CH4
conversion and H2 yield, and they are defined as:
XCH4

YH2

nCO nCO2
FCH4

QH
FCH4

(9)

(10)

The predicted reactor performances from both models were


illustrated in Fig. 2. When very few membrane surface area
was installed, i.e., Cep below 15 Km for the operation conditions specified, the predicted methane conversion and
hydrogen yield from both models are very close, with the
predictions from PFR model slightly higher than those predicted from CSTR model. Under such conditions, the hydrogen
yield increases almost linearly with increasing the membrane
permeation capacity, indicating that the separation is the
controlling step for hydrogen production. When Cep varies
from 15 to 60 Km, the predicted methane conversion and H2
yield from PFR model are obviously higher than those from the
CSTR model. When membrane permeation capacity is high
enough, i.e., Cep higher than 60 Km for the condition studied,
the reactor performances from both models became close
again. This is because that the plug flow reactor can provide
a higher driving force for hydrogen permeation through
membranes than a CSTR. However, if very few membrane
surface areas are employed, the effect of reactor hydrodynamics on reactor performances is minimized. When
membrane surface area is large enough, the equilibrium of the
reactions became the controlling step, and the influence of
membrane permeation capacity on reactor performance is
also suppressed.

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 1 7 9 8 e1 1 8 0 9

1.2

4.0

1.1

3.5

1.0

3.0

0.9

2.5

0.8

2.0

CH4 conversion CSTR

0.7

1.5

CH4 conversion PFR


H2 yield CSTR

0.6

1.0

H2 yield PFR

0.5
0.4

0.20e0.28 mm) palladium membrane tubes, each has outside


diameter of 4.7 mm were installed in the reactor. Table 1
compares experimental data of methane conversion,
hydrogen production rate from the membrane, and product
gas composition with the predictions from the current model,
as well as the predictions from the PFR model of Ye et al. [15].
The permeation efficiency h is taken as 0.39, as suggested by
Adris et al. [12]. It can be seen from the table that except for
some points (for example CH4 concentration at 640  C and CO2
concentration at 542  C), the predictions are in satisfactory
agreement with the experimental data in general. The
discrepancies could be caused by several factors, for example,
the error on selection of the membrane permeation efficiency
factor, errors on reactor temperature, pressure and gas
composition samplings, chemical reactions in the reactor
freeboard that is not considered in the current model, etc. It is
also worth mentioning that the predictions from both models
are almost identical. This is consistent with the previous
statement that when Cep is low, the predictions from both
CSTR and PFR models are very close.
Roy [39] tested a fluidized bed membrane reactor with high
flux membrane tubes. The tubes had a substrate thickness of
76 mm and palladium coating of 5.2 mm on both the outside and
inside surfaces. The outer diameter of the tubes was
3.175 mm. The following parameters were evaluated by
experimentation: the permeability pre-exponential factor k
was 7.85  109 mol (m s Pa0.72); activation energy Ep was
11.5 kJ/mol; pressure exponent M was 0.72. Hence these
numbers were employed in the current model for this case.
Nine high flux membrane tubes were installed in a 100 mm
diameter reactor, where oxygen was fed to provide the heat. In
the fluidized bed reactor, the coverage of membrane outside
surface by catalyst dust, exposure of membrane tubes to two
different phases (bubble phase and dense phase) and the gas

Hydrogen yield ( mol/mol CH4 )

CH4 conversion ( - )

11802

0.5

20

40

60

80

100

120

140

160

0.0
180

Membrane permeation capacity (Km)

Fig. 2 e Influence of Cep on methane conversion and


hydrogen yield. (FCH4 [ 226.41 mol/h, FS [ 0; SC [ 3.0,
OC [ 0, T [ 650  C, P [ 2.0 MPa, PM [ 0.1 MPa).

3.
Comparison of model predictions with
experimental data
The FBMR process has been intensively studied by researchers
in recent years. Some experimental data are available from
the literature for model validations. Experimental results by
Adris et al. [12], Roy [39], and Mahecha-Botero et al. [5] are used
to validate the current model performances.
Adris et al. [12] reported experimental results from a pilotscale fluidized bed membrane reactor for reaction temperatures from 447 to 640  C and steam to carbon molar ratio (SC)
of 2.4. The reactor has a diameter of 97 mm and length of
1.143 m. Twelve thin-walled (nominal wall thickness of

Table 1 e Comparison of equilibrium model predictions with experimental data of Adris et al. [12]. (FCH4 [ 74.2 mol/h,
SC [ 2.4, OC [ 0, Fs [ 80 mol/h, Cep [ 0.4 Km, P [ 0.98 MPa, PM [ 0.4 MPa, h [ 0.39).
Bed temperature ( C)

447

494

542

594

640

Methane conversion ()

Experimental data
CSTR prediction
PFR prediction

0.12
0.11
0.11

0.17
0.16
0.16

0.24
0.22
0.22

0.33
0.32
0.34

0.43
0.41
0.42

H2 production (mol/h)

Experimental data
CSTR prediction
PFR prediction

1.70
1.73
1.77

2.50
2.51
2.55

3.57
3.50
3.61

4.81
4.77
4.95

6.23
6.23
6.30

Experimental data
CSTR prediction
PFR prediction
Experimental data
CSTR prediction
PFR prediction
Experimental data
CSTR prediction
PFR prediction
Experimental data
CSTR prediction
PFR prediction

61.7
62.1
62.7
0.10
0.19
0.30
9.5
7.8
7.6
28.7
29.4
29.4

49.6
51.9
52.3
0.40
0.52
0.50
11.5
9.6
9.5
38.5
38.0
37.6

37.6
41.8
42.3
1.20
1.25
1.24
13.0
11.2
10.8
48.2
45.8
45.9

27.3
31.8
32.1
3.20
2.73
2.60
13.5
12.1
11.7
56.0
53.4
55.3

19.5
23.9
24.0
5.60
4.80
4.70
13.3
12.0
11.7
61.6
59.3
59.5

Product gas composition


(vol%, dry basis)
CH4

CO

CO2

H2

11803

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 1 7 9 8 e1 1 8 0 9

Table 2 e Comparison of model predictions with experimental data of Roy [39]. (FH2 O [ 138.33 mol/h, FS [ 45 mol/h,
Cep [ 8.16 Km, P [ 0.68 MPa, PM [ 0.14 MPa, k [ 7.85 3 10L9 mol/(m.s Pa0.72), Ep [ 11.5 KJ/mol, M [ 0.72).
Bed P (MPa)

0.68
0.68
0.68
0.68
0.68
0.68
0.68
0.68
0.99
0.99
0.99
0.68
0.68
0.68
0.68
0.68
0.78
0.88
0.99

Bed T ( C)

650
650
650
650
650
650
650
650
600
625
650
576
600
625
649
600
600
600
600

SC

4.1
4.1
4.1
4.1
4.1
4.1
3.1
2.4
3.1
3.1
3.1
4.1
4.1
4.1
4.1
4.1
4.1
4.1
4.1

OC

0.44
0.45
0.5
0.56
0.62
0.35
0.35
0.35
0.40
0.40
0.40
0.45
0.45
0.45
0.45
0.45
0.45
0.45
0.45

CH4 conversion ()

Hydrogen production (mol/h)

Experimental

Predicted
h 0.3

Predicted
h 0.1

Experimental

Predicted
h 0.3

Predicted
h 0.1

0.76
0.77
0.79
0.81
0.82
0.74
0.66
0.59
0.52
0.56
0.6
0.62
0.67
0.73
0.77
0.68
0.67
0.63
0.62

0.92
0.93
0.94
0.95
0.96
0.91
0.85
0.77
0.74
0.80
0.86
0.78
0.84
0.89
0.93
0.84
0.84
0.83
0.85

0.87
0.86
0.88
0.90
0.92
0.85
0.77
0.70
0.63
0.69
0.75
0.70
0.76
0.82
0.87
0.76
0.86
0.85
0.85

18.6
18.6
18.5
17.6
17.1
19.1
20.5
21.4
19.0
21.0
23.2
14.6
15.9
17.5
18.6
14.8
16.0
16.3
17

43.0
42.6
41.4
39.8
38.1
45.1
50.9
54.8
40.2
45.2
49.9
32.3
36.1
39.7
42.5
36.1
39.2
42.1
44.9

21.7
21.6
21.2
20.6
19.9
22.5
24.1
25.1
23.5
26.6
29.7
15.8
17.8
19.8
21.5
17.8
23.5
25.2
27.06

phase mass transfer resistance may affect the membrane tube


permeability [39]. They used permeation efficiency h to
account for the influence of above factors on the membrane
permeability. However, the actual value of h was not disclosed. In the current model, values of 0.3 and 0.1 were tried.
Table 2 gives results of model predictions in comparison with
experimental results. It can be seen that the predictions from
h 0.1 was closer to the experimental results.
Mahecha-Botero et al. [5] tested a pilot-scale fluidized bed
membrane reactor for the production of hydrogen. The
reactor was operated under steam methane reforming and
autothermal reforming conditions, without membranes and
with membranes of different total areas. Heat was added
either externally or via direct air addition. The reactor is
contained in a stainless steel vessel of height 2 m and rectangular cross-sectional area of 48.4 cm2. The vessel can hold
up to six double-sided membrane modules with a nominal
permeation area of 300 cm2. The membrane modules contained Pd/Ag foil of thickness 25 mm sealed onto a porous
metal backing with a barrier layer to prevent interdiffusion.
The authors suggested in another paper [3] that the
membrane efficiencies were ranging from w0.6 to 0.9. Here
the efficiency of 0.6 is applied in the current model. Table 3
compares the experimental results with the model predictions. It can be seen that the model slightly over-estimated the
natural gas conversion, the hydrogen concentration in the
reactor off gas and hydrogen production rate.

4.
Effects of operation conditions on carbon
formation
Coke formation is a serious operational problem in conventional steam reforming. The presence of the solid carbon can

deactivate the catalyst and, hence degrading the performance


of the reforming system [14]. The operating regime that the
solid carbon may appear should be avoided in practical
applications.

4.1.
Influence of steam to carbon ratio and membrane
permeation capacity
Fig. 3 shows the influence of steam to carbon ratio and
membrane permeation capacity on carbon formation under
the operation conditions of FCH4 1000 mol/h, FS 0, OC 0.2,
P 2.0 MPa, T 650  C, PM 0.1 MPa. It can be seen that carbon
formation is suppressed by higher steam to carbon ratio. The
in-situ removal of hydrogen through membranes enhanced
the carbon formation. The coking boundary moves toward
higher SC values as the reactor Cep is raised. For example, for
the operation conditions investigated, when Cep increased
from 1.0 Km to 10.0 Km, the coking boundary increased from
an SC value of 0.8e1.0. With higher membrane permeation
capacities per unit feed of methane and steam, more
hydrogen is removed from the reactor, and reaction (R8) is
pushed forward to produce more solid carbon.

4.2.
Influence of temperature and membrane permeation
capacity
Fig. 4 shows the influence of temperature and membrane
permeation capacity on carbon formation for the operation
conditions of FCH4 1000 mol/h, FS 0, OC 0.2, P 2.0 MPa,
T 650  C, PM 0.1 MPa. The range of the temperature
investigated is between 450 and 850  C. It should be pointed
out that when palladium alloy membrane is installed, the
reactor temperature is usually maintained below 700  C for
the protection of membrane. It can be seen that temperature
is a strong factor for the formation of solid carbon. The

11804

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 1 7 9 8 e1 1 8 0 9

Carbon yield (mol/mol CH4)

21.3
47.7
54.3
40.6
45.3
29.4
72.2
74.2
50.9
37.5
47.6
N/A
N/A
N/A
13.8
39.3
42.0
41.5
41.1
N/A
49.6
48.2
36.6
32.6
39.7

Prediction
Expt

Permeate H2 flow (mol/h)

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km

0.35
0.30
0.25
0.20
0.15
0.10

Prediction

53.7
51.0
31.4
49.2
43.5
39.2
35.6
30.8
26.8
19.9
19.9
14.2
11.2
15.5

Expt

43.5
33.3
22.2
40.1
41.1
33.3
29.4
20.6
22.0
17.5
16.2
11.9
8.7
8.7

0.00
0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

Steam to carbon ratio (mol steam/ mol CH 4)

amount of carbon initially increased and then decreased with


the temperature increasing. Again, the membrane enhanced
the carbon formation. The coking boundary moves toward
lower temperatures as the reactor Cep is raised. For example,
for the operation conditions investigated, if Cep increased from
1.0 Km to 10 Km, the coking boundary decreased from 600  C
to 500  C. For the formation of solid carbon, reaction (R5) is
endothermic and (R9eR11) are exothermic. With the temperature increasing, reactions (R9eR11) are shifted to the right
direction while reaction (R8) went to the opposite direction.
The net effect of temperature on carbon formation behaves
then as shown in Fig. 4.

4.3.
Influence of reactor pressure and membrane
permeation capacity
Reactions (R8eR11) are related to solid carbon formation in the
reforming process. In these reactions, reaction (R8) should be
0.30

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km

0.25

Carbon yield (mol/mol CH4 )

a In bed gas sampling not performed. The methane conversion was calculated from reactor off gas composition analysis.

0.30
0.27
0.45
0.39
0.50
0.51
0.66
0.69
0.58
0.78
0.78
0.92
0.96
0.95
0.29
0.22
0.40
0.25a
0.66
0.66
0.43a
0.67
0.46a
0.73
0.69
0.81
0.76a
0.73
1
2
3
4
5
6
7
8
9
10
11
12
13
14

0
0
0
0.09
0.09
0.09
0.09
0.09
0.18
0.18
0.18
0.18
0.18
0.18

0
0
0.35
0
0
0
0
0
0.35
0.35
0.35
0.35
0.35
0

40
40
40
40
40
40
20
20
40
40
40
20
13.3
13.3

0.75
1.00
1.00
0.75
0.75
1.00
0.75
1.00
1.00
1.00
1.00
1.00
1.00
1.00

N/A
N/A
N/A
0.10
0.03
0.03
0.03
0.03
0.10
0.03
0.03
0.03
0.03
0.03

Prediction

Fig. 3 e Influence of steam to carbon ratio and membrane


permeation capacity on carbon yield. (FCH4 [ 1000 mol/h,
FS [ 0, OC [ 0.2, P [ 2.0 MPa, T [ 650  C, PM [ 0.1 MPa).

Expt

Membrane
area (m2)

OC

Natural gas
feed
(mol/h)

Reactor
pressure
(MPa)

Permeate
pressure
(MPa)

Methane conversion ()

H2 molar fraction in ROG (dry) (%)

0.05

Expt

Table 3 e Comparison of model predictions with experimental data of Mahecha-Botero et al. [5]. (Fs [ 0, SC [ 3.0, Cep [ 0.4 Km, P [ 0.98 MPa, k [ 1.37 3 10L3 mol/
(m$h$Pa0.5), Ep [ 20.5 KJ/mol, M [ 0.5, h [ 0.6).

0.40

0.20

0.15

0.10

0.05

0.00
450

500

550

600

650

700

750

800

850

Temperature ( C )

Fig. 4 e Influence of temperature and membrane


permeation capacity on carbon yield. (FCH4 [ 1000 mol/h,
FS [ 0, SC [ 0.8, OC [ 0.2, P [ 2.0 MPa, PM [ 0.1 MPa).

11805

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 1 7 9 8 e1 1 8 0 9

250

0.15

0.10

200

0.0008

150

0.0006

100

50

0
400

0.5

1.0

1.5

2.0

2.5

3.0

500

550

600

650

700

750

800

0.0000
850

the dominant process as it is the source of solid carbon. These


processes follow LeChateliers principle, so that the formation
of solid carbon benefit from low pressure. In an FBMR system,
a perm-selective membrane installed in the fluidized bed
reduces the adverse effect of high pressure by removing
product hydrogen. Ye et al. [15] concluded from their simulation that over the range of conditions investigated, at low
permeation capacity, methane conversion decreases with
increasing reactor pressure. At medium permeation capacity,
the influence of pressure on methane conversion is almost
neutral. At high permeation capacity, methane conversion
increases with increasing reactor pressure. Here for the
formation of solid carbon, the same trends from the influence
of membrane permeation capacity and pressure were
observed as that of the methane conversion, as shown in
Fig. 5. At low permeation capacity (Cep 0.4e5.0 Km), carbon

0.30

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km
Cep=20.0 Km

0.25

0.20

Temperature ( C )

3.5

Fig. 5 e Influence of pressure and membrane permeation


capacity on carbon yield. (FCH4 [ 1000 mol/h, FS [ 0,
OC [ 0.2, SC [ 1.0, T [ 650  C, PM [ 1.0 MPa).

Fig. 7 e Influence of temperature and membrane


permeation capacity NH3 molar fraction and yield.
(FCH4 [ 1000 mol/h, SC [ 3.0, OC [ 0.2, P [ 2 0.0 MPa,
PM [ 0.1 MPa).

formation decreases with increasing reactor pressure. At


medium permeation capacity (Cep 10.0 Km), the influence of
pressure on carbon formation is almost neutral. If a high
permeation capacity (Cep 20.0 Km) could be used, carbon
formation would increase with increasing reactor pressure.

4.4.
Influence of oxygen to carbon ratio and membrane
permeation capacity
Obviously, oxygen input can reduce the amount of solid
carbon formation through reaction (R10). Again, the in-situ
removal of product hydrogen could enhance the carbon
formation. Fig. 6 shows the influence of OC and membrane
permeation capacity on carbon formation at the operation
condition of FCH4 1000 mol/h, FS 0, SC 0.8, T 650  C,
P 2.0 MPa, PM 0.1 MPa. It can be seen that with increasing
the oxygen to carbon ratio, the mole of carbon formed is
decreasing for all Cep values. The coking boundary moves
toward higher OC ratio as the reactor Cep is raised. For the
operation conditions investigated, when Cep increased from
180
160

0.15

0.10

0.05

0.1

0.2

0.3

0.4

0.5

Oxygen to carbon ratio (mol O2 / mol CH4 )

Fig. 6 e Influence of oxygen to carbon ratio and membrane


permeation capacity on carbon yield. (FCH4 [ 1000 mol/h,
FS [ 0, OC [ 0.2, SC [ 0.8, T [ 650  C, P [ 2 MPa,
PM [ 0.1 MPa).

Molar fraction of NH3 (ppm)

Carbon yield (mol/mol CH4 )

450

0.0002

Pressure (MPa)

0.00
0.0

0.0004

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km

0.05

0.00
0.0

0.0010

NH3 yield (mol/mol CH4 )

0.20

0.0012

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km

140

0.0009
0.0008

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km

0.0007
0.0006

120

0.0005
0.0004

100

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km

80
60
40
0.0

0.5

1.0

1.5

2.0

2.5

3.0

0.0003
0.0002

NH 3 yield (mol/mol CH4 )

0.25

300

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km
Cep=20.0 Km

Molar fraction of NH3 (ppm)

Carbon yield (mol/mol CH4)

0.30

0.0001
0.0000
3.5

Pressure (MPa)

Fig. 8 e Influence of pressure and membrane permeation


capacity on NH3 molar fraction and yield. (FCH4 [ 1000 mol/
h, SC [ 3.0, OC [ 0.2, T [ 650  C, PM [ 0.1 MPa).

11806

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 1 7 9 8 e1 1 8 0 9

0.0010

120

0.0006
110

0.0005
0.0004

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km

90
0

120

Molar fraction of NH 3 (ppm)

0.0008
0.0007

100

0.0011

0.0009

NH3 yield (mol/mol CH 4 )

Molar fraction of NH3 (ppm)

130

0.0012

130

0.0010
0.0009

110

0.0008
100

0.0007
0.0006

90

0.0003

70
4

0.0002

0.0

0.1

Steam to carbon ratio (mol steam/ mol CH4)

0.4 Km to 20 Km, the coking boundary for OC increased from


0.15 to 0.50.

5.
Effects of operation conditions on NH3
formation

Molar fraction (-)

0.26

0.20
0.18
0.16
0.14
0.12

the reactant gas is below 10 ppm. Hence the formation of


these species will not be discussed here, and only NH3
formation is presented.

Reaction (R12) is exothermic; therefore increasing operation


temperature can shift the equilibrium backward and suppress
NH3 formation. With the in-situ removal hydrogen with
membranes, the hydrogen concentration in the reactant gases
is decreased. Hence the membrane helps to reduce the
formation of NH3 i.e., higher membrane permeation capacity
helps to reduce the NH3 concentration in the system. This is
confirmed by the model simulation results as shown in Fig. 7.

0.20
0.16
0.12

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km

0.08
0.04

0.10
1.0

1.5

2.0

2.5

3.0

3.5

4.0

0.0002

0.24

0.22

0.5

0.5

0.28

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0Km
Cep=10.0 Km

0.24

0.08
0.0

0.4

0.0003

Fig. 11 e Influence of steam to carbon ratio and membrane


permeation capacity on NH3 molar fraction and yield.
(FCH4 [ 1000 mol/h, FS [ 0, SC [ 3.0, T [ 650  C,
P [ 2.0 MPa, PM [ 0.1 MPa).

Molar fraction of H2 and N2

0.28

0.3

0.0004

5.1.
Influence of temperature and membrane permeation
capacity

When air is introduced into the reactor for autothermal


operation, the nitrogen in the air could involve with reactions
to generate NO, NO2, N2O, NO2, N2H4 and NH3 through reactions (R12eR17), which are not desired. The formation of these
by-products was investigated for the conditions of
FCH4 1000 mol/h, FS 0, SC 0.5e5.0, OC 0.05e0.5,
T 400e800  C, P 0.5e3.0 MPa, PM 0.1 MPa,
Cep 0.4e10.0 Km. It was found that under such operation
conditions, the concentration of NO, NO2, N2O, NO2, N2H4 in

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km

0.2

0.0005

Oxygen to carbon ratio (mol O2/ mol CH4)

Fig. 9 e Influence of steam to carbon ratio and membrane


permeation capacity on NH3 molar fraction and yield.
(FCH4 [ 1000 mol/h, FS [ 0, OC [ 0.2, T [ 650  C,
P [ 2.0 MPa, PM [ 0.1 Mpa).

0.30

Cep=0.4 Km
Cep= 1.0 Km
Cep= 5.0 Km
Cep=10.0 Km

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km

80

NH3 yield (mol/mol CH4 )

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km

140

4.5

5.0

5.5

Steam to carbon ratio (mol steam / mol CH4 )

Fig. 10 e Influence of steam to carbon ratio and membrane


permeation capacity on N2 and H2 molar fraction.
(FCH4 [ 1000 mol/h, FS [ 0, OC [ 0.2, T [ 650  C,
P [ 2.0 MPa, PM [ 0.1 Mpa) (solid symbols: N2, empty
symbols: H2).

0.0

0.1

0.2

0.3

Cep=0.4 Km
Cep=1.0 Km
Cep=5.0 Km
Cep=10.0 Km
0.4

0.5

Steam to carbon ratio (mol steam/ mol CH4)

Fig. 12 e Influence of oxygen to carbon ratio and


membrane permeation capacity on N2 and H2 molar
fractions. (FCH4 [ 1000 mol/h, FS [ 0, SC [ 3.0, T [ 650  C,
P [ 2.0 MPa, PM [ 0.1 MPa) (solid symbols: N2, empty
symbols: H2).

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 1 7 9 8 e1 1 8 0 9

It can be seen that for the operation conditions studied, the


NH3 concentration in the product gas is between 50 and
220 ppm and the NH3 yield is around 3  105 mol/mol of CH4.

11807

with increasing the oxygen to carbon ratio, the N2 concentration increases and H2 concentration decreases. The overall
tendency of NH3 formation by the influence of oxygen to
carbon ratio then behaves as illustrated in Fig. 11.

5.2.
Influence of pressure and membrane permeation
capacity
Fig. 8 shows the influence of pressure and membrane
permeation capacity on the NH3 formation. It can be seen that
the NH3 concentration increases with increasing the operation pressure. This can be explained by LeChateliers principle
that NH3 formation is favored by high pressure as the product
moles are less than the reactants in reaction (R12). Again
membrane helps to reduce the formation of NH3.

5.3.

Influence of steam to carbon ratio

When other operation conditions are fixed, the steam to


carbon ratio affects the NH3 formation by the following
processes:
A) Higher steam feed diluted the N2 concentration in the
reactant streams, which favors the NH3 formation;
B) At lower steam to carbon ratios, when it is increased, the
methane conversion is promoted, and lead to higher H2
concentration in the reactant streams. If steam to carbon
ratio is further increased, the extra steam dilutes the
hydrogen in the stream.
Fig. 9 shows the influence of steam to carbon ratio on NH3
molar fraction and yield for Cep 0.4, 1.0, 5.0, 10.0 Km. It can be
seen that NH3 yield increases with increasing the steam to
carbon ratio. However, its molar concentration reached
maximum values around steam to carbon ratio of 1.0e1.5.
Fig. 10 shows the N2 and H2 concentration in the reactor.
When steam to carbon ratio increases, the N2 concentration is
decreasing, while the H2 concentration increase until SC
reached approximately 2.5, and then decreases with further
increasing of the SC values.

5.4.

Influence of oxygen to carbon ratio

When other operation conditions are fixed, the oxygen to


carbon ratio affects the NH3 formation by the following two
processes:
A) With higher oxygen to carbon ratio, the N2 concentration
in the reactant stream is increased, which favors the
formation of the NH3;
B) Higher air feed to the reactor consumed and diluted the H2
concentration in the reactant streams which in return
decreased molar number of NH3 in the stream.
Fig. 11 shows the influence of oxygen to carbon ratio on
NH3 molar fraction and yield for Cep 0.4, 1.0, 5.0, 10.0 Km. It
can be seen that NH3 yield increases with increasing the
oxygen to carbon ratio. However, its molar concentration
reached maximum values around oxygen to carbon ratio of
0.35. Fig. 12 shows the N2 and H2 concentration variations in
the reactor with the steam to carbon ratio. It can be seen that

6.

Conclusions

A reaction/separation coupled thermodynamic equilibrium


model was developed to simulate hydrogen production with
fluidized bed membrane reactors from steam methane
reforming reactions with or without air addition. The model
assumes a CSTR for bed hydrodynamics. The model considers
CH4, H2O, O2 and N2 in the feed stream and the reactor off gas
constitutes of 15 species, including CO, CO2, H2O, CH4, H2, O2,
N2, C2H2, C2H4, and NO, NO2, N2O, NH3, N2H4 and solid carbon.
Model predictions on reactor performance show satisfactory
agreement with the experimental results of Adris et al. [12],
Roy [39], and Mahecha-Botero et al. [5].
Parametric studies of the carbon formation from the steam
methane reforming with the model show that:
1. Over the range of condition investigated, the in-situ removal
of hydrogen through membranes enhances the carbon
formation. Carbon formation is suppressed by higher steam
to carbon ratio.
2. Temperature is a strong factor for the formation of solid
carbon. Over the range of conditions studied, the amount of
carbon formed initially increased and then decreased with
the temperature increasing. Solid carbon yield decreases
with increasing steam to carbon ratio.
3. For the conditions studied, at low permeation capacity
(Cep 0.4e5.0 Km), carbon formation decreases with
increasing reactor pressure. At medium permeation
capacity (Cep 10.0 Km), the influence of pressure on
carbon formation is almost neutral. If a high permeation
capacity (Cep 20.0 Km) could be used, carbon would
increase with increasing reactor pressure.
4. O2 input can reduce the amount of solid carbon formation.
NH3 formation was also studied by the current model.
Parametric studies of the NH3 formation from the steam
methane reforming with the model show that over the range
of operations studied:
1. Increasing operation temperature can shift the equilibrium
backward and decrease NH3 formation.
2. The membrane helps to reduce the formation of NH3 i.e.,
higher membrane permeation capacity helps to reduce the
NH3 concentration in the system.
3. NH3 concentration and yield increase with increasing the
operation pressure.
4. NH3 yield increases with increasing steam to carbon ratio.
However, its concentration reached maximum values
around steam to carbon ratio of 1.0e1.5.
5. NH3 yield increases with increasing oxygen to carbon ratio.
However, its concentration reached maximum values
around steam to carbon ratio of 0.30e0.35.

11808

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 1 7 9 8 e1 1 8 0 9

Acknowledgement
Financial support from the National High Technology Research
and Development Program of China (2009AA05Z102) and the
Fundamental Research Funds for the Central Universities
(project # 2009ZZ0013) are gratefully acknowledged.

Notation

aik
Ak
Cep
EP
FCH4
Fs
Gt
Goi
Gi
GC
QH
k
m
ni
M
OC
P
P0
PRH
PMH
PM
R
SC
T
VC
XCH4
YH2
yi
h

number of atoms of the kth element present in each


molecule of species i in reacting system, e
total molar flow rate of kth element in the feed,
mol h1
membrane permeation capacity (membrane surface
area/thickness), Km
activation energy, J mol1
molar flow rate of methane feed, mol h1
molar flow rate of sweep gas, mol h1
total Gibbs free energy, J mol1
Gibbs free energy of species i at its standard state,
J mol1
Gibbs free energy of pure species i at operating
conditions, J mol1
molar Gibbs free energy of solid carbon, J mol1
total molar flow rate of hydrogen extracted through
perm-selective membranes, mol h1
pre-exponential factor, mol Km1 h1 MPaM
number of sub-separators used in the model of Ye
et al. [15], e
molar flow rate of specie i in the product gas and
solid carbon, mol h1
pressure exponent in Equation (1)
oxygen to air ratio, e
total absolute pressure of the reactor, MPa
standard state pressure, MPa
absolute partial pressure of hydrogen in the reactor,
MPa
absolute partial pressure of hydrogen in the
permeate side, MPa
absolute pressure in the permeate side, MPa
universal gas constant, J mol1 K1
steam to carbon ratio, e
temperature,  C
molar volume of solid carbon, m3 mol1
conversion of CH4, defined in Equation (9), e
H2 yield, defined in Equation (10), e
molar fraction of species i in the product gas, e
permeation efficiency factor, e

references

[1] Xie D, Grace JR, Lim CJ. Experimental study of gas and solid
circulation in an internally circulating fluidized bed
membrane reactor cold model. Chem Eng Sci 2009;64:
2599e606.

[2] Grace JR, Elnashaie SSEH, Lim CJ. Hydrogen production in


fluidized beds with in-situ membranes. Int J Chem Reactor
Eng 2005;3:A41.
[3] Mahecha-Botero A, Chen Z, Grace JR, Elnashaie SSEH, Lim CJ.
Comparison of fluidized bed flow regimes for steam methane
reforming in membrane reactors: a simulation study. Chem
Eng Sci 2009;64:3598e613.
[4] Chen Z, Grace JR, Lim CJ, Li A. Experimental studies of pure
hydrogen production in a commercialized fluidized bed
membrane reactor with SMR and ATR catalysts. Int J
Hydrogen Energy 2007;32:2359e66.
[5] Mahecha-Botero A, Boyd T, Gulamhusein A, Comyn N,
Lim CJ, Grace JR, et al. Pure hydrogen generation in
a fluidized-bed membrane reactor: experimental findings.
Chem Eng Sci 2008;63:2752e62.
[6] Chen Z, Po F, Grace JR, Lim CJ, Elnashaie S, Mahecha-Botero A,
et al. Sorbent-enhanced/membrane-assisted steam-methane
reforming. Chem Eng Sci 2008;63:170e82.
[7] Prasad P, Elnashaic SSEH. Novel circulating fluidized-bed
membrane reformer for the efficient production of ultraclean
fuels from hydrocarbons. Ind Eng Chem Res 2002;41:6518e27.
[8] Xu JG, Froment GF. Methane steam reforming, methanation
and wateregas shift: I. Intrinsic kinetics. AIChE J 1989;35:
88e96.
[9] Jin W, Gu X, Li S, Huang P, Xu N, Shi J. Experimental and
simulation study on a catalyst packed tubular dense
membrane reactor for partial oxidation of methane to
syngas. Chem Eng Sci 2000;55:2617e24.
[10] Snoeck JW, Froment GF, Fowlest M. Kinetic study of the
carbon filament formation by methane cracking on a nickel
catalyst. J Catal 1997;169:250e62.
[11] Tottrup PB. Kinetics of decomposition of carbon monoxide
on a supported nickel catalyst. J Catal 1976;42:29e36.
[12] Adris AM, Lim CJ, Grace JR. The fluidized-bed membrane
reactor for steam methane reforming: model verification and
parametric study. Chem Eng Sci 1997;52:1609e22.
[13] Xie D. Hydrogen production from reforming or partial
oxidation of natural gas. In: High temperature processes in
chemical engineering. Vienna: Process Eng Engineering
GmbH; 2010.
[14] Grace JR, Li X, Lim CJ. Equilibrium modeling of catalytic
steam reforming of methane in membrane reactors with
oxygen addition. Catal Today 2001;64:141e9.
[15] Ye G, Xie D, Qiao W, Grace JR, Lim CJ. Modeling of fluidized
bed membrane reactors for hydrogen production from steam
methane reforming with Aspen Plus. Int. J. Hydrogen Energy
2009;34:4755e62.
[16] Sieverts A, Zapf G. Solubility of H and D in solid PD(I). Z Phys
Chem A 1935;17:359e64.
[17] Federico G, Erik EE, Ma YH. Effects of surface activity, defects
and mass transfer on hydrogen permeance and n-value in
composite palladium e porous stainless steel membrane.
Catal Today 2006;118:24e31.
[18] Nam SE, Lee SH, Lee KH. Preparation of a palladium alloy
composite membrane supported in a porous stainless steel
by vacuum electrodeposition. J Membr Sci 1999;153:163e73.
[19] McCool BA, Lin YS. Nanostructured thin palladiumesilver
membranes: effects of grain size on gas permeation
properties. Mater J Sci 2001;36:3221e7.
[20] Ward TL, Dao T. Model of hydrogen permeation behavior in
palladium membrane. J Membr Sci 1999;153:211e31.
[21] Unemoto A, Kaimai A, Sato K, Otake T, Yashiro K, Mizusaki J,
et al. Surface reaction of hydrogen on a palladium alloy
membrane under co-existence of H2O, CO, CO2 or CH4. Int. J.
Hydrogen Energy 2007;32:4023e9.
[22] Xie D, Adris AM, Lim CJ, Grace JR. Test on a modular fluidized
bed membrane reactor for autothermal steam methane
reforming. Acta Energiae Solaris Sinica 2009;30:704e7.

i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 5 ( 2 0 1 0 ) 1 1 7 9 8 e1 1 8 0 9

[23] Knovel. Knovel critical tables. Online version available at.


2nd ed, http://www.knovel.com/knovel2/Toc.jsp?
BookID761&VerticalID0; 2003 [accessed 25.08.08].
[24] Smith WR, Missen RW. Chemical reaction equilibrium
analysis: theory and algorithms. New York: Wiley; 1982.
[25] Van ZF, Storey SH. The computation of chemical equilibra.
Cambridge: Cambridge University Press; 1970.
[26] Holub R, Vonka P. The chemical equilibrium of gaseous
systems. Dordrecht: Reidel; 1976.
[27] McDonald CM, Floudas CA. Global optimization for the phase
and chemical equilibrium problem: application to the NRTL
equation. Comput Chem Eng 1995;19:1111e39.
[28] Nichita DV, Gomez S, Luna E. Multiphase equilibria calculation
by direct minimization of Gibbs free energy with a global
optimization method. Comput Chem Eng 2002;26:1703e24.
[29] Li XT, Grace JR, Lim CJ, Watkinsona AP, Chen HP, Kim JR.
Biomass gasification in a circulating fluidized bed. Biomass
and Bioenergy 2004;26:171e93.
[30] Li XT, Grace JR, Watkinson AP, Lim CJ, Ergudenler A.
Equilibrium modeling of gasification: a free energy
minimization approach and its application to a circulating
fluidized bed coal gasifier. Fuel 2001;80:195e207.
[31] Jarungthammachote S, Dutta A. Equilibrium modeling of
gasification: Gibbs free energy minimization approach and
its application to spouted bed and spout-fluid bed gasifiers.
Energy Conversion Manage 2008;49:1345e56.

11809

[32] Lu Y, Guo L, Zhang X, Yan Q. Thermodynamic modeling and


analysis of biomass gasification for hydrogen production in
supercritical water. Chem Eng J 2007;131:233e44.
[33] Li Y, Wang Y, Zhang X, Mi Z. Thermodynamic analysis of
autothermal steam and CO2 reforming of methane. Int J
Hydrogen Energy 2008;33:2507e14.
[34] Chan SH, Wang HM. Thermodynamic analysis of natural-gas
fuel processing for fuel cell applications. Int. J. Hydrogen
Energy 2000;25:441e9.
[35] Rossi CCRS, Alonso CG, Antunes OAC, Guirardello R,
Cardozo-Filho L. Thermodynamic analysis of steam
reforming of ethanol and glycerine for hydrogen production.
Int J Hydrogen Energy 2009;34:323e32.
[36] Amin NAS, Yaw TC. Thermodynamic equilibrium analysis
of combined carbon dioxide reforming with partial
oxidation of methane to syngas. Int. J. Hydrogen Energy
2007;32:1789e98.
[37] Semelsberge TA, Borup RL. Thermodynamic equilibrium
calculations of hydrogen production from the combined
processes of dimethyl ether steam reforming and partial
oxidation. J Power Sources 2006;155:340e52.
[38] Chase Jr MW, Davies CA, Downey Jr JR, Frurip DJ,
McDonald RA, Syverud AN. JANAF thermochemical tables.
J. Phys. Chem. Ref. Data 1985;14(Suppl. 1). 3rd ed.
[39] Roy S. FBMR with high-flux membrane and O2 input. PhD
thesis. Alberta: University of Calgary; 1998.

You might also like