You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/316439180

Methane Dry Reforming over Ni-Co/Al2O3 : Kinetic Modelling in a Catalytic


Fixed-bed Reactor

Article  in  International Journal of Chemical Reactor Engineering · April 2017


DOI: 10.1515/ijcre-2016-0170

CITATIONS READS

2 308

4 authors, including:

Yacine BENGUERBA Mirella Virginie


Ferhat Abbas University of Setif Université des Sciences et Technologies de Lille 1
36 PUBLICATIONS   96 CITATIONS    25 PUBLICATIONS   431 CITATIONS   

SEE PROFILE SEE PROFILE

Barbara Ernst
University of Strasbourg
33 PUBLICATIONS   751 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Hydrogen separation View project

Molecular Dynamics Simulation View project

All content following this page was uploaded by Yacine BENGUERBA on 28 January 2019.

The user has requested enhancement of the downloaded file.


DE GRUYTER International Journal of Chemical Reactor Engineering. 2017;

Yacine Benguerba1,2 / Mirella Virginie3,4 / Christine Dumas3,4 / Barbara Ernst3,4

Methane Dry Reforming over Ni-Co/Al2 O3 :


Kinetic Modelling in a Catalytic Fixed-bed
Reactor
1
Department of Processes Engineering, Université Ferhat Abbas, Sétif-1, 19000 Sétif, Algeria, E-mail:
benguerbayacine@yahoo.fr
2
Laboratoire de Génie des Procédés Chimiques, Université Ferhat Abbas, Sétif-1, 19000 Sétif, Algeria, E-mail:
benguerbayacine@yahoo.fr
3
Université de Strasbourg, IPHC, RePSeM, 25 rue Becquerel, 67087 Strasbourg, France
4
CNRS, UMR7178, 67087 Strasbourg, France

Abstract:
The dry reforming of CH4 was investigated in a catalytic fixed-bed reactor to produce hydrogen at different
temperatures over supported bimetallic Ni-Co catalyst. The reactor model for the dry reforming of methane
used a set of kinetic models: The Zhang et al model for the dry reforming of methane (DRM); the Richardson-
Paripatyadar model for the reverse water gas shift (RWGS); and the Snoeck et al kinetics for the coke-deposition
and gasification reactions. The effect of temperatures on the performance of the reactor was studied. The
amount of each species consumed or/and produced were calculated and compared with the experimental
determined ones. It was showed that the set of kinetic model used in this work gave a good fit and accurately
predict the experimental observed profiles from the fixed bed reactor. It was found that reaction-4 and reaction-
5 could be neglected which could explain the fact that this catalyst coked rapidly comparatively with other
catalyst. The use of large amount of Ni-Co will lead to carbon deposition and so to the catalyst deactivation.
Keywords: dry reforming of methane, greenhouse gas, modelling, catalytic reactor, coke deposition
DOI: 10.1515/ijcre-2016-0170

1 Introduction
The CO2 methane reforming has been, for a long time, a subject of several studies. This process has important
environmental and economic implications, since it utilizes both methane and carbon dioxide, two greenhouse
gases that cause global warming, to produce useful chemical products (Al-Fatesh, Fakeeha, and Abasaeed 2011).
In recent years, the Ni-based catalysts have gained much attention despite their deactivation due to the
excessive deposition of carbon during the catalytic conversion process. For this purpose, many investigators
(Hu and Ruckenstein 2002; Bradford and Vannice 1999) concentrated their studies on the suppression of carbon
formation which is stopping the catalytic reforming of methane to be used commercially.
The supported group VIII metal catalysts have been used extensively in the methane dry reforming (DRM).
The noble metals (e. g. Ru, Rh, Pd, Pt, Ir) as well as non-noble metals (e. g. Ni, Co, Fe) were found to be catalyti-
cally active towards this reaction (Rostrup-Nielsen and Bak Hasen 1993). The noble metal is active and have been
found to be less prone to coke formation under reforming conditions. The drawback of noble metal catalysts is
their high cost and limited availability. In contrast, non-noble metals are less expensive and widely available.
However, dry reforming being strongly endothermic requires high temperatures (typically 800–900 °C) and this
leads to the rapid deactivation of the catalyst by carbon deposition and/or metal sintering for non-noble (Liu
2010).
At the industrial level, the catalyst properties and operating conditions must be carefully selected to mini-
mize undesirable transient modifications of the active catalytic phase, e. g., by deactivation and coking (Chen
et al. 2010). Several methods and different nickel-based catalysts have been proposed for reducing coke forma-
tion, however, a solution has still not been found (Corthals et al. 2011; Tomishige and Fujimoto 1998; Zhou et
al. 2011).
It has been reported that the replacement of 50 mol% of the NiO by CoO in NiO-MgO (Ni/Mg=1), improved
the performance of the catalyst in the methane to syngas conversion process; and drastically reduced the carbon

Yacine Benguerba is the corresponding author.


© 2017 Walter de Gruyter GmbH, Berlin/Boston.

Authenticated | benguerbayacine@yahoo.fr author's copy


Download Date | 4/24/17 6:13 PM
Benguerba et al. DE GRUYTER

formation on the catalyst (Choudhary and Mamman 1998). The addition of Co leads to the strong adsorption
capacity of CO2 which, in turn, favours the elimination of carbon. There are reports that the catalyst activity and
stability can be improved through formation of a homogeneous alloy between Co and Ni (Takanabe et al. 2005).
Xiaohong et al. (2010), Zhang, Wang, and Dalai (2007), and Koh et al. (2007), reported that bimetallic catalysts
performed better than the corresponding monometallic catalysts. Zhang, Wang, and Dalai (2008) showed that
the change of the metal dispersion and metal particle size facilitated the improved activity and coke suppression
of Ni–Co bimetallic catalysts. In addition, in their previous work, Zhang, Wang, and Dalai (2007) showed that
reducing Ni and Co content from 6.1 and 9.3 to 3.6 and 4.9 mol% (metal base), respectively, rendered to eliminate
carbon deposition.

CH4 + CO2 ⟷ 2CO + 2H2 , ΔH0298 = +247kJmol−1 [1]

CO2 + H2 ⟷ CO + H2 O, ΔH0298 = +41.7kJmol−1 [2]

The major drawback of DRM is the high temperatures required to reach high conversion. This is due to the
highly endothermic nature of the reaction.

CH4 ⟷ C + 2H2 , ΔH0298 = +74.87kJmol−1 [3]

C + H2 O → CO + H2 , ΔH0298 = +131.325kJmol−1 [4]

C + CO2 ⟷ 2CO, ΔH0298 = +172kJmol−1 [5]

The present work deals with Ni-Co alumina supported catalyst (16.5 wt.% Ni; 16.5 wt.% Co). The objective
was to show the effect of promoting the Ni catalyst with Co on the catalytic activity and the resistance to coke for-
mation during the reaction. Previous results (Benguerba et al. 2015) showed that a Ni/Al2 O3 catalyst (33 wt.%
Ni content) yielded similar results. The experimental work was conducted to validate the one-dimensional
mathematical model for the catalytic fixed-bed reactor operating at steady-state conditions.

2 Experimental apparatus and operative conditions


The experimental device contained a stainless-steel module (ID = 0.02 m; L = 0.25 m) in which the quartz
tube (ID = 0.0085 m; L = 0.22 m) is placed, an oven composed of THERMOCOAX temperature resistant heat-
ing wire with a WEST 820 temperature regulator, and an on-line AGILENT Technologies M200 gas micro-
chromatograph.
The tube was fed with a standard mixture consisting of CH4 : CO2: N2 = 1:1:8 (molar ratio) regulated by
BROOKS 5850 TR mass flow controllers.
In the module of tubular geometry, the feed gas flowed inside the tube (Figure 1), which was filled with a
catalyst composed of nickel-cobalt supported on alumina (200 mg of Ni(16.5 %)-Co(16.5 %)/Al2 O3 ), prepared
by precipitation of the metal oxalates. The catalyst was diluted with glass beads. The measurements were
made based on a temperature range of 723–973K. The feed (10 % CH4 , 10 % CO2 , and 80 % N2 ) flow rate was
50.3 NmL/min (Figure 1).

Authenticated | benguerbayacine@yahoo.fr author's copy


Download Date | 4/24/17 6:13 PM
DE GRUYTER Benguerba et al.

Figure 1: The catalytic fixed-bed membrane reactor module.

2.1 Reactor model


In the simulation of the dry reforming of methane into synthesis gas in a fixed-bed reactor, a one-dimensional
homogeneous model is applied (Froment and Bischoff 1990). Temperature and concentration gradients are
accounted for in axial direction only. The reactor model for the DRM used the kinetics of Zhang, Wang, and
Dalai (2009) for a bimetallic Ni-Co/Al-Mg-O catalyst coupled with the Richardson-Paripatyadar (Richardson
and Paripatyadar 1990) kinetic for the RWGS. Coke formation in the reforming of hydrocarbon is known to
cause the catalyst deactivation (Zhang, Wang, and Dalai 2007; Koh et al. 2007; Zhang, Wang, and Dalai 2008;
2009; Richardson and Paripatyadar 1990; Olsbye, Wurzel, and Mleczko 1997). The coke is expected to be formed
mainly via the methane decomposition (3) and gasified by steam (4) and by CO2 (5), which are favourable under
CO2 -reforming conditions (Rostrup-Nielsen 1983; Claridge et al. 1993; Snoeck, Froment, and Fowles 1997a;
1997b; 2002).
The reactor model consists of the following set of ordinary differential equations:
The continuity equations for the various components are
dXCH4 ρb ωL
= (r1 + r3 ) [6]
dz F0CH
4

dXCO2 ρb ωL
= (r1 + r2 + r5 ) [7]
dz F0CO
2

dXCO ρ ωL
= b0 (2∗r1 + r2 + r4 + 2𝑟5 ) [8]
dz FCH
4

dXH2 ρb ωL
= (2∗r1 − r2 + 2∗r3 + 𝑟4 ) [9]
dz F0CH
4

dXH2 O ρb ωL
= (r2 − r4 ) [10]
dz F0CH
4

The energy equation is


5
dT ρ L 4Uw L
= b ∑ ri (−ΔH)i − (T − Tw ) [11]
dz uρg Cp i=1 Dt uρg Cp

Authenticated | benguerbayacine@yahoo.fr author's copy


Download Date | 4/24/17 6:13 PM
Benguerba et al. DE GRUYTER

The pressure drop is considered negligible.


The integration of the first-order differential eqs 6–11 in axial direction was performed by means of a fourth-
order Runge-Kutta routine. The rates (r1 , r2 , r3 , r4 and r5 ) were estimated using expressions given in Table 1. The
Thermodynamic and rate constants are given in Table 2. To solve the differential equations, the initial values for
the species mole fractions, as well as the temperature of the feed at the entry of the reactor (z = 0) are given in
Table 3. These values are used to calculate the partial pressures used to calculate the rates (ri , i = 1:5). By Solving
differential eqs 6–11), the new conversions/yields and temperature for the next step (z+Δz) are calculated. This
procedure is repeated up to reaching the length L of the reactor.

Table 1: Reaction rates equations: DRM; RWGS; coke formation and gasification.
Dry reforming of methane

2
u�1 PCH4 PCO2 (PCO PH2 )
Reaction-1 (Zhang, Wang, and Dalai 2009) r1 = (u�u�u�2 ,1 PCO2 +KCH4 ,1 PCH4 )
(1 − KP1 (PCH4 PCO2 )
)
k2 KCO2 ,2 KH2 ,2 PCO2 PH2 (PCO PH2 O )
Reaction-2 (Richardson and Paripatyadar 1990) r2 = 2 (1 − KP2 (PCO2 PH2 )
)
(1+KCO2 ,2 PCO2 +KH2 ,2 PH2 )

Coke formation and gasification reactions

P2
H ⎞
k3 KCH4 ,3 ⎛
⎜PCH4 − K 2 ⎟
⎜ ⎟
p3
Reaction-3 (Snoeck, Froment, and Fowles 1997a) r3 = ⎝ ⎠
2
(1+KCH4 ,3 PCH4 + K 1 P1.5 )
H2 ,3 H2
k4 PH O P
KH O,4
( P 2 − KCO )
Reaction-4 (Snoeck, Froment, and Fowles 2002)
H2 P4
r4 = 2
PH O 2
(1+KCH4 ,4 PCH4 + K 1 2 + 1 P1.5 )
KH ,4 H2
H2 O,4 PH2 2
k5 PCO P
KCO,5 KCO ,5
( P 2 − KCO )
Reaction-5 (Snoeck, Froment, and Fowles 2002)
CO P5
r5 = 2
PCO 2
1 2)
(1+KCO,5 PCO + K
CO,5 KCO2 ,5 PCO

Table 2: Thermodynamic and rate constants. Constants are in [mol, kg, s, bar].
Parameter Value Parameter Value
k1 (Zhang, Wang, and 1.3510 −8
exp (− 3115.22
T
) KH2O,4 (Snoeck, Froment, 4.7310−6 exp (+ 97770
RT
)
Dalai 2009) and Fowles 2002)
k2 (Richardson and 0.35106 exp (− 81030
RT
) KCH4,4 (Snoeck, Froment, 0
3.49exp (− RT )
Paripatyadar 1990) and Fowles 2002)
k3 (Snoeck, Froment, and 6.95103 exp (− 58893
RT
) KH2,4 (Snoeck, Froment, and 1.8310+13 exp (− 216145
RT
)
Fowles 1997a) Fowles 2002)
k4 (Snoeck, Froment, and 5.55109 exp (− 166397
RT
) KCO,5 (Snoeck, Froment, 7.3410−6 exp (+ 100395
RT
)
Fowles 2002) and Fowles 2002)
k5 (Snoeck, Froment, and 1.341015 exp (− 243835
RT
) KCO2,5 (Snoeck, Froment, 2.8110+7 exp (− 104085
RT
)
Fowles 2002) and Fowles 2002)
KCO2,1 (Zhang, Wang, and 9.2510−8 exp (+ 4883.32
T
) Kp1 (Zhang, Wang, and 6.7810+14 exp (− 259660
RT
)
Dalai 2009) Dalai 2009)
KCH4,1 (Zhang, Wang, and 2.4610−7 exp (+ 4606.68
T
) Kp2 (Richardson and 56.4971exp (− 36580
RT
)
Dalai 2009) Paripatyadar 1990)
KCO2,2 (Richardson and 0.5771exp (+ 9262
RT
) Kp3 (Snoeck, Froment, and 2.9810+5 exp (− 84400
RT
)
Paripatyadar 1990) Fowles 1997a)
KH2,2 (Richardson and 1.494exp (+ 6025
RT
) Kp4 (Snoeck, Froment, and 1.382710+7 exp (− 125916
RT
)
Paripatyadar 1990) Fowles 2002)
KCH4,3 (Snoeck, Froment, 0.21exp (− 567
RT
) Kp5 (Snoeck, Froment, and 1.939310+9 exp (− 168527
RT
)
and Fowles 1997a) Fowles 2002)
KH2,3 (Snoeck, Froment, and 5.1810+7 exp (− 133210
RT
)
Fowles 1997a)

Table 3: Operating conditions and reactor dimensions.

Authenticated | benguerbayacine@yahoo.fr author's copy


Download Date | 4/24/17 6:13 PM
DE GRUYTER Benguerba et al.

Inlet pressure, Pt 1 bar


Inlet temperature, T0 25 °C
Total inlet flow rate, Q0 50.3 Nml/min
CH4 /CO2 – feed ratio 1:1
CH4 /N2 – feed ratio 1:8
Inlet mole fraction of CH4 , 9.94 %
Inlet mole fraction of CO2 , 10.33 %
Reactor length, L 0.22 m
Reactor diameter, Dt 0.008 m
Wall temperature, Tw 450-700 °C
Global heat transfer coefficient, U 200 Watt/m2
Mean catalyst particle diameter 0.32 mm

Table 4: CH4 and CO2 conversions at different temperatures.


Temperature CH4 Conversion (%) – Simulated CH4 Conversion CO2 Conversion CO2 Conversion
(°C) (%) – (%) – Simulated (%) –
Experimental Experimental
450 0.2210 0.2121 0.1709 0.1595
500 0.4380 0.4172 0.3024 0.2767
550 0.6967 0.6937 0.4339 0.4110
600 0.8572 0.8245 0.5683 0.5854
650 0.9137 0.9108 0.7369 0.7398
700 0.9435 0.9583 0.8742 0.8513

2.2 Kinetics

For the DRM and the RWGS reactions, the kinetics given by Zhang, Wang, and Dalai (2009)and by Richardson-
Paripatyadar (Richardson and Paripatyadar 1990) respectively were used. The kinetic models developed by
Snoeck, Froment, and Fowles (1997a, 1997b, 2002) were used for the carbon deposition and gasification reac-
tions (Table 1). For the Zhang et al kinetic model, we have added a term to the original model to account for the
reverse reaction (1).
Thermodynamic and rate constants for the rate equations for DRM, RWGS and for coke formation and
gasification are given in Table 2.

3 Model validation – results and discussion


The simulations of the DRM in the fixed-bed catalytic reactor was performed under operation conditions given
in Table 3.
Figure 2summarizes all the results obtained numerically and experimentally. It shows that the set of kinetic
rates could be used with confidence in the case of dry reforming of methane over a Ni-Co/Al2 O3 catalyst. The
CH4 , CO2 molar flows and consequently conversions (Table 4) are correctly determined using the numerical
model whereas the calculated CO and H2 molar flows are slightly different from experimental ones. This low
difference between the calculated and experimental CO and H2 molar flow profiles is probably due to the
misleading RWGS kinetic model estimation. Among different models which are not cited here, it was found
that Richardson-Paripatyadar RWGS-kinetic model (Richardson and Paripatyadar 1990) gives the best fit.

Authenticated | benguerbayacine@yahoo.fr author's copy


Download Date | 4/24/17 6:13 PM
Benguerba et al. DE GRUYTER

Figure 2: Comparison between the experimental and the simulated molar fluxes obtained at the exit of the reactor at op-
erating conditions (Table 3); open circle CH4 -experimental; dotted line CH4 -simulated; open square CO2 -experimental;
line CO2 -simulated; open triangle CO-experimental; dashed line CO-simulated; open diamond H2 -experimental; long
dash line H2 -simulated.

Figure 3 shows, the influence of the temperature, on the rates of CO2 reforming of methane over the
Ni-Co/Al2 O3 catalyst at atmospheric pressure. The competition between reaction-1 and reaction-3 to convert
CH4 to syngas is clearly presented, showing that the dominant reaction is reaction-1. Reaction-3 is lower, but
coke is significantly produced.

Figure 3: Effect of the wall temperature on the dry reforming reactions rates at operating conditions (Table 3); dotted line
reaction_1; medium dash line reaction_2; short dash line reaction_3; continuous line reaction_4; broken line reaction_5.

Unlike the case of Ni/Al2 O3 (Benguerba et al. 2015), the use of a bimetallic catalyst, Ni-Co/Al2 O3 makes
the reaction-4 very low or even negligible and reaction-5 very low. This is probably due to the high content of

Authenticated | benguerbayacine@yahoo.fr author's copy


Download Date | 4/24/17 6:13 PM
DE GRUYTER Benguerba et al.

nickel and cobalt in the catalyst giving greater metal sizes and consequently made the catalyst less coke resistant
(Zhang, Wang, and Dalai 2008). This finding was confirmed experimentally because this catalyst was found to
coke very quickly after a few hours of operation: The large quantities of coke produced by reaction-3 are not
gasified by the very low reaction-4 and reaction-5 leading to a coked catalyst. This result is in full agreement
with the results obtained experimentally by Zhang, Wang, and Dalai (2008): the higher is the Ni-Co content
(16.5 wt.% Ni; 16.5 wt.% Co), the higher is the carbon deposition extent. For temperatures 650 and 700 °C, the
reaction of gasification by CO2 is very significant at the entrance of the catalytic bed but it soon becomes very low
in the center and near the exit of the catalyst bed. Figure 3 shows that the peak of production of the reaction-3
(carbon deposition) starts at Z > 0.5 for low temperatures (T = 450 and 500 °C), but begins to move toward the
inlet of catalyst bed (Z = 0.48) for high temperatures. This peak of the reaction-3 is accompanied by a very small
gasification reaction-4 and reaction-5. It is concluded that the deactivation of the catalyst should be mainly in
the center of the catalyst and a little less at the exit of the catalyst bed.
Figure 4 shows the simulated temperature profiles at different wall temperatures (Tw ). The obtained results
indicated the global endothermicity of the chemical scheme. In the entrance of the catalyst bed, it is observed
a deep decrease of the temperature (endothermic global reaction) especially at high wall temperatures. As the
wall temperature increased, the endothermicity of the global reaction increased. This is due to the activation of
the reaction set in the endothermic way.

Figure 4: Evolution of the reactor temperature profiles at operating conditions (Table 3) with different wall temperatures;
broken line 450 °C; long dash line 500 °C; medium dash line 550 °C; short dash line 600 °C; dotted line 650 °C; continuous
line 700 °C.

The problem that must be pointed out is the heat transfer efficiency from the external wall to the catalyst
bed. The consumed heat in the endothermic process was not compensated in the catalyst zone by the heat
transmitted from the external wall. This raises several questions about the effectiveness of heat transfer in such
reactors. The heat transfer limitation for packed-bed reactors could alternatively be solved by coating a thin
layer of the catalyst on the inner walls of the reactor, thus avoiding the pressure drop associated with packed
bed reactors. The temperature could be easily controlled and the reactor will operate near isothermally.

4 Conclusion
Dry reforming of methane over a Ni-Co/Al2 O3 catalyst has been kinetically modelled. The reaction scheme
consisted of five reactions: the DRM, the RWGS, methane decomposition, and the two gasification reactions.
The kinetic rates used in this study was the Zhang et al model for the DRM; the Richardson-Paripatyadar model
for the RWGS reaction; and the Snoeck et al models for coke deposition and gasification reactions. It was found
that reaction-4 and reaction-5 could be neglected which could explain the fact that this catalyst coked rapidly

Authenticated | benguerbayacine@yahoo.fr author's copy


Download Date | 4/24/17 6:13 PM
Benguerba et al. DE GRUYTER

comparatively with other catalyst. The use of large amount of Ni-Co will lead to carbon deposition and so to
the catalyst deactivation.
The simulation of the wall-heated reactor system showed that the conversion, as well as the temperature
profiles, are dependent on the heated-wall temperature. It is noticed a weak heat transfer efficiency which is a
major drawback of such reactors.

Nomenclature
cp heat capacity (J kg−1 K−1 )
Dt inlet reactor diameter (m)
𝐹u�0 initial molar flow for species i (mol/s)
k reaction rate constant (see Table 2)
K adsorption constant (see Table 2)
Kp equilibrium constant for reaction i (see Table 2)
L reactor length (m)
P pressure (bar)
r specific rate of reaction (mol kg−1 s−1 )
R universal gas constant (J mol−1 K−1 )
T temperature (K)
u gas velocity (m s−1 )
Uw overall heat transfer coefficient (W/m2 K)
Xi conversion for (CH4 and CO2 ); Yield for (CO, H2 O, H2 )
z axial coordinate (m)
𝜌u� catalyst bed density (kg/m3 )
𝜌u� gas density (kg/m3 )
ω catalyst mass (kg)
𝜀 bed porosity (-)
ΔH heat of reaction (kJ mol−1 )

References
Al-Fatesh, A. S. A., A. H. Fakeeha, and A. E. Abasaeed. 2011. “E昀fects of Promoters on Methane Dry Reforming over Ni Catalyst on a Mixed
(Α-Al2 o3 +Tio2 -P25) Support .” International Journal of the Physical Sciences 6 (36): 8083–8092.
Benguerba, Y., L. Dehimi, M. Virginie, C. Dumas, and B. Ernst. 2015. “Modelling of Methane Dry Reforming over Ni/Al2 O3 Catalyst in a fixed-
Bed Catalytic Reactor .” Reaction Kinetics, Mechanisms, and Catalysis 114 (1): 109–119.
Bradford, M. C. J., and M. A. Vannice. 1999. “CO2 Reforming of CH4 .” Catalysis Reviews – Science and Engineering 41 (1): 1–42.
Chen, D., R. Lodeng, H. Svendsen, and A. Holmen. 2010. “Hierarchical Multiscale Modeling of Methane Steam Reforming Reactions .” Indus-
trial & Engineering Chemistry Research 50: 2600–2612.
Choudhary, V. R., and A. S. Mamman. 1998. “Simultaneous Oxidative Conversion and CO2 or Steam Reforming of Methane to Syngas over
Coo-Nio-Mgo Catalyst .” Journal of Chemical Technology and Biotechnology 73: 345–350.
Claridge, J. B., M. L. H. Green, S. C. Tsang, A. P. E. York, A. T. Ashcro昀t, and P. D. Battle. 1993. “A Study of Carbon Deposition on Catalysts during
the Partial Oxidation of Methane to Synthesis Gas .” Catalysis Letters 22: 299–305.
Corthals, S., T. Witvrouwen, P. Jacobs, and B. Sels. 2011. “Development of Dry Reforming Catalysts at Elevated Pressure: D-Optimal Vs. Full
Factorial Design .” Catalysis Today 159: 12–24.
Froment, G. F., and K. B. Bischo昀f. 1990. Chemical Reactor Analysis and Design., 2nd. New York, NY: Wiley.
Hu, Y. H., and E. Ruckenstein. 2002. “Binary Mgo-Based Solid Solution Catalyst for Methane Conversion to Syngas .” Catalysis Reviews – Sci-
ence and Engineering 44 (3): 423–453.
Koh, A. C. W., L. W. Chen, W. Keeleong, B. F. G. Johnson, T. Khimyak, and J. Y. Lin. 2007. “Hydrogen or Synthesis Gas Production via the Partial
Oxidation of Methane over Supported Nickel-Cobalt Catalysts .” International Journal of Hydrogen Energy 32: 725–730.
Liu, D., et al. 2010. “A Comparative Study on Catalyst Deactivation of Nickel and Cobalt Incorporated MCM-41 Catalysts Modified by Plat-
inum in Methane Reforming with Carbon Dioxide .” Catalysis Today 154 (3): 229–236.

Authenticated | benguerbayacine@yahoo.fr author's copy


Download Date | 4/24/17 6:13 PM
DE GRUYTER Benguerba et al.

Olsbye, U., T. Wurzel, and L. Mleczko. 1997. “Kinetic and Reaction Engineering Studies of Dry Reforming of Methane over a Ni/La/Al2 O3 Cata-
lyst .” Industrial & Engineering Chemistry Research 36 (12): 5180–5188.
Richardson, J. T., and S. A. Paripatyadar. 1990. “Carbon Dioxide Reforming of Methane with Supported Rhodium .” Applied Catalysis 61:
293–309.
Rostrup-Nielsen, J. R 1983. “Catalytic Steam Reforming.” In Catalysis Science and Technology., edited by J. R. Anderson, and M. Boudart, Vol. 5,
1–118. Berlin: Springer.
Rostrup-Nielsen, J. R., and J.-H. Bak Hasen. 1993. “CO2-Reforming of Methane over Transition Metals .” Journal of Catalysis 144: 38–49.
Snoeck, J.-W., G. F. Froment, and M. Fowles. 1997a. “Filamentous Carbon Formation and Gasification: Thermodynamics, Driving Force, Nucle-
ation, and Steady-State Growth .” Journal of Catalysis 169: 240–249.
Snoeck, J.-W., G. F. Froment, and M. Fowles. 1997b. “Kinetic Study of the Carbon Filament Formation by Methane Cracking on a Nickel Cata-
lyst .” Journal of Catalysis 169: 250–262.
Snoeck, J.-W., G. F. Froment, and M. Fowles. 2002. “Steam/CO2 Reforming of Methane. Carbon Filament Formation by the Boudouard Reac-
tion and Gasification by CO2 , by H2 , and by Steam: Kinetic Study .” Industrial & Engineering Chemistry Research 41: 4252–4265.
Takanabe, K., K. Nagaoka, K. Nariai, and K. I. Aika. 2005. “Titania Supported Cobalt and Nickel Bimetallic Catalysts for Carbon Dioxide Re-
forming of Methane .” Journal of Catalysis 232: 268–275.
Tomishige, K., and K. Fujimoto. 1998. “Ultra-Stable Ni Catalysts for Methane Reforming by Carbon Dioxide .” Catalysis Surveys from Asia 2:
3–15.
Xiaohong, L. I., A. I. Jun, L. I. Wenying, and L. I. Dongxiong. 2010. “Ni-Co Bimetallic Catalyst for CH4 Reforming with CO2 .” Frontiers of Chemical
Science and Engineering 4: 476–480.
Zhang, J. G., H. Wang, and A. K. Dalai. 2007. “Development of Stable Bimetallic Catalysts for Carbon Dioxide Reforming of Methane .” Jour-
nal of Catalysis 249: 300–310.
Zhang, J. G., H. Wang, and A. K. Dalai. 2008. “E昀fects of Metal Content on Activity and Stability of Ni-Co Bimetallic Catalysts for CO2 Reform-
ing of CH4 .” Applied Catalysis 339: 121–129.
Zhang, J. G., H. Wang, and A. K. Dalai. 2009. “Kinetic Studies of Carbon Dioxide Reforming of Methane over Ni-Co/Al-Mg-O Bimetallic Cata-
lyst .” Industrial & Engineering Chemistry Research 48: 677–684.
Zhou, C., L. Zhang, A. Swiderski, W. Yang, and W. Blasiak. 2011. “Study and Development of a High Temperature Process of Multi-
Reformation of CH4 with CO2 for Remediation of Greenhouse Gas .” Energy 36: 5450–5459.

Authenticated | benguerbayacine@yahoo.fr author's copy


Download Date | 4/24/17 6:13 PM

View publication stats

You might also like