Вы находитесь на странице: 1из 32

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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

Available at www.sciencedirect.com

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

Review

Review of methane catalytic cracking for hydrogen production

Ashraf M. Amin a, Eric Croiset b,*, William Epling b


a
Chemical Engineering & Pilot Plant Department, National Research Center, Dokki, Giza, Egypt
b
Chemical Engineering Department, University of Waterloo, 200 University Avenue West, Waterloo, Ontario, Canada N2L 3G1

article info abstract

Article history: Methane catalytic cracking is a process by which carbon monoxide-free hydrogen can be
Received 12 September 2010 produced. Despite the fact that hydrogen produced from methane cracking is a pure form
Received in revised form of hydrogen, methane cracking is not used on an industrial scale for producing hydrogen
5 November 2010 since it is not economically competitive with other hydrogen production processes.
Accepted 9 November 2010 However, pure hydrogen demand is increasing annually either in amount or in number of
Available online 28 December 2010 applications that require carbon monoxide-free hydrogen. Currently, hydrogen is produced
primarily via catalytic steam reforming, partial oxidation, and auto-thermal reforming
Keywords: of natural gas. Although these processes are mature technologies, CO is formed as
Hydrogen a by-product, and in order to eliminate it from the hydrogen stream, complicated and
Methane cracking costly separation processes are required. To improve the methane catalytic cracking
Fluidized bed economics, extensive research to improve different process parameters is required. Using a
Carbon nanotubes highly active and stable catalyst, optimizing the operating conditions, and developing
Catalyst suitable reactors are among the different areas that need to be addressed in methane
cracking. In this paper, catalysts that can be used for methane cracking, and their deac-
tivation and regeneration are discussed. Also, methane catalytic cracking kinetics
including carbon filament formation, the reaction mechanisms, and the models available
in the literature for predicting reaction rates are presented. Finally, the application of
fluidized beds for methane catalytic cracking is discussed.
ª 2010 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved.

1. Introduction and overview applications such as in automobiles and power stations [1].
However, hydrogen nowadays is mostly produced via fossil
The majority of current energy needs is supplied by combus- fuel reforming and fossil fuels are expected to remain a major
tion of non-renewable energy sources, namely fossil fuels, hydrogen source in the long term [2]. Presently, hydrogen
and is associated with the release of large quantities of finds application as a chemical compound rather than a fuel in
greenhouse gases (GHG), especially carbon dioxide (CO2), and commercial operations, but if hydrogen is used to replace
other harmful emissions to the atmosphere. The gradual existing fuels, appropriate methods for large-scale production
depletion of these fossil fuel reserves, and efforts to combat must be developed.
pollution and greenhouse gas emissions, have generated Hydrogen is the simplest element and the most plentiful
considerable interest in using alternative sources of energy. gas in the universe. Yet, hydrogen never occurs by itself in
Hydrogen gas may one day replace fossil fuels for various nature; it is found combined with other elements such as

* Corresponding author. Tel.: þ1 5198884567x36472; fax: þ1 5197464979.


E-mail address: ecroiset@uwaterloo.ca (E. Croiset).
0360-3199/$ e see front matter ª 2010 Professor T. Nejat Veziroglu. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2010.11.035
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2905

oxygen and carbon, i.e. water or hydrocarbons, so these to the lower energy demand for methane catalytic cracking
substances must be decomposed/reformed to get H2 [3]. To compared to steam reforming, there is no need for additional
release hydrogen, heat can be applied to hydrocarbons and energy for steam generation or gas treatment. The heat
water to break down the molecules (thermo-chemical), requirement for catalytic cracking can be covered by burning
and electric charging (electrolysis) or a photolytic process can w15e20% of the hydrogen produced, which further reduces
be used to decompose water. Some separation processes CO2 emissions [16].
combine heat and electricity (steam electrolysis). Bacteria and Thermal methane cracking is not feasible at moderate
algae can also be used to produce H2 from biomass. temperatures. To achieve a reasonable yield, a temperature
Currently, the main process for producing hydrogen is higher than 1200  C is required [3,17]. Supported metal cata-
steam reforming of natural gas [1]. Steam reforming is lysts can be used to catalytically decompose hydrocarbons to
a multiple stage process. The first stage is the highly endo- produce hydrogen at more moderate temperatures. Nickel is
thermic catalytic reforming of methane (DH0298 ¼ 206 kJ/mol particularly active for hydrocarbon cracking, especially for
CH4) conducted at high temperature (800e900  C), while the methane [18]. Cobalt and iron can also be used to catalyze
second stage is the catalytic water gas shift (WGS) reaction, methane cracking but their carbon/active site capacities are
occurring in two steps: the first stage is conducted at much lower than that of nickel, with additional problems in
(400e500  C) to reduce the carbon monoxide (CO) concentra- the case of cobalt, which are associated with its higher cost
tion to 2e5%, and the second stage is conducted at (177e257  C) and toxicity [19]. Metal catalysts are usually deposited on
to reduce CO to 1%. The third stage is the separation of the supports such as SiO2 or Al2O3 and the performance of the
H2eCO2 mixture using pressure-swing adsorption (PSA) [4e6]. catalyst depends to some extent on the combination of metal
This process needs other auxiliary steps, such as a desulphu- and support [20].
rization unit and a steam generation section. In addition to Early methane cracking was done in fire-brick furnaces at
steam reforming, partial oxidation is also used to generate 1500  C to produce soot [2]. In 1966, the HYPRO process was
hydrogen from fossil fuels [7], but the produced hydrogen is introduced and was the first commercial process for the
still mixed with CO and CO2, which again needs a complicated catalytic cracking of methane and gaseous hydrocarbons for
separation process as in the steam reforming case. hydrogen production [21]. The HYPRO process was based on
Increasing demand for CO-free hydrogen has increased a circulating fluidized bed, operating at temperatures up to
interest in the direct catalytic cracking of natural gas [8,9], 980  C and at atmospheric pressure. The carbon produced
described by Equation (1). The two reaction products are during the cracking step was burned in a regenerator to supply
hydrogen and carbon, the latter being essentially in the form the process energy requirements and to regenerate the cata-
of filamentous carbon or carbon nanotubes [10]. lyst [2]. The HYPRO process could convert a dry gas mixture of
methane and light hydrocarbons to 90% hydrogen with 10%
CH4 4Cs þ 2H2 ; DH025  C ¼ 74:8 kJ=mol (1) unconverted methane, using a 7% Ni/Al2O3 catalyst. The
capital cost of this process was lower than the steam
As a result of methane cracking, only hydrogen is produced reforming process [2,21]. However, the operating cost of the
as a gaseous product in a mixture with unreacted methane. process was high, partially due to the circulation of catalyst
Separation of methane and hydrogen can be achieved easily between the two fluidized beds, which required a certain
by absorption or membrane separation to produce a stream of pressure drop be maintained between the different reactors
99% by volume hydrogen, which is much simpler than the [22]. More recent economic studies of the methane cracking
need for further complicated separation processes that deal process indicate that the viability of the process depends in
with CO2 or CO [2,7]. This would be particularly important in a large part on the selling price of the filamentous carbon [23].
the case of proton-exchange membrane (PEM) fuel cell appli- The major challenge in a continuous process for hydrogen
cations, since the Pt-based electrocatalyst is poisoned by CO production using methane catalytic cracking is the regenera-
[11]. More specifically, the CO concentration in hydrogen tion of the spent catalyst [24], which is also critical in the
streams used as fuel for PEM fuel cells must be lower than overall economics of the process. During filamentous carbon
20 ppm in order to prevent significant deactivation [5]. The formation in methane catalytic cracking the active nickel
carbon nanotubes produced as a solid product are a commer- atom is found on the tip of the filament [18,25], as shown in
cially valuable material; they are useful in many applications, Fig. 1. Being located on the tip, the catalytic activity of the
especially in adsorption processes and catalysis or as an supported catalyst can be maintained for an extended period
option for storing carbon [2,12]. of time. However, the catalyst deactivates due to the forma-
Unlike the steam reforming process, the catalytic decom- tion of encapsulating carbon that ultimately blocks access to
position of methane (CDM) does not include water gas shift and the nickel. Steam gasification or air oxidation can be used to
preferential oxidation of CO, which considerably simplifies the regenerate the deactivated catalyst [26].
process and may reduce the hydrogen production costs [2,13]. Different reactors for continuous methane cracking have
The energy required for methane catalytic cracking is nearly been proposed, such as a set of parallel fixed-bed reactors
one half that required for steam reforming per mole of methane alternating between different conditions or a fluidized bed/
decomposed (for steam reforming ΔH0298 ¼ þ253.2 kJ/mol, regenerator combination [27]. In either system, the ability of
for methane cracking ΔH0298 ¼ þ74.8 kJ/mol) [2,9,14,15]. Per the catalyst to accumulate significant amounts of filamentous
mole of hydrogen produced, the energy requirement is carbon prior to its deactivation allows for substantial catalyst
37.4 kJ/mol H2 in methane catalytic cracking compared to time-on-stream or residence times in the reactor or reaction
63.3 kJ/mol H2 in the steam reforming process [15]. In addition zone. For fluidized bed systems, regeneration of the catalyst
2906 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

where csol
CH4 =H2 is the solubility of carbon in the active metal (e.g.
Ni) in contact with the hydrogen/methane mixture (units of
mol C/volume of active metal particle).
Equilibrium methane conversion, as well as the number of
moles of CH4, H2 and C (for an initial 100 moles of CH4) as
a function of temperature is shown in Fig. 2. Equilibrium
conversion increases with increasing temperature, starting
from w30% conversion at 500  C to almost complete conver-
sion at 1000  C.
The value of the equilibrium constant can be predicted
from Gibbs free energy equations and temperature dependent
equations are available in the literature, all of which consider
carbon as graphite [13e34]. For example, Villacampa et al. [13]
Fig. 1 e Metal particle on the tip of a carbon filament
proposed the following correlation for the Gibbs free energy:
(reprinted from [18], with permission from Elsevier).
DG0 ðJ=molÞ ¼ 89; 658:88  102:27T  0:00428T2

requires continuous catalyst circulation between the cracker  2; 499; 358:99=T (4)
and regenerator sections. Another advantage, coincident with Ginsburg et al. [34] developed another equation for calcu-
this continuous operation, is that a fluidized bed system lating the Gibbs free energy for methane cracking:
allows continuous addition or removal of solids as the reac-
tion proceeds. Furthermore, in a fluidized bed reactor the bed DG0 ðJ=molÞ ¼ 58; 886:7 þ 270:55T þ 0:0311T2  3  106 T3
of catalyst particles behaves as a continuously stirred tank þ 291; 405:7=T  54:598T lnðTÞ (5)
reactor, which improves heat and mass transfer [27], pre-

Both Equations (4) and (5) lead to a zero DG at 547 C, indi- 0
venting radial and axial temperature gradients and providing
cating that, thermodynamically, methane decomposes to
better utilization of the catalyst particles, due to the high
form carbon and hydrogen at temperatures above 547  C.
external mass transfer rates. Also, small catalyst particles can
However, methane cracking has been reported in the litera-
be used while maintaining small pressure drop.
ture at temperatures as low as 500  C [24]. This shows that
This review will discuss different aspects related to
calculating the Gibbs free energy using graphite may not be
methane catalytic cracking, subdivided into three parts; the
adequate. Indeed, the carbon nanotubes produced during
first part discusses the fundamentals of the thermodynamics,
methane cracking have chemical and thermodynamic prop-
catalyst development, carbon filaments formation, and cata-
erties different from those of graphite [25].
lyst deactivation associated with methane catalytic cracking.
Rostrup-Nielsen [28] studied the deposition of carbon on
The second part includes different methods used for catalyst
various nickel catalysts during methane cracking and found
regeneration, and reaction mechanisms and reaction rate
a deviation from the equilibrium data when assuming that
models. The last part discusses the engineering of the process,
carbon is deposited as graphite. They attributed the deviation,
focusing on application of fluidized bed reactors, and scale up
between the experimental data and the equilibrium calcula-
from laboratory to industrial scale.
tion taking into consideration graphite, to the surface disorder
of the formed carbon structure and the higher surface energy
of the carbon filaments. The authors proposed an equation to
2. Thermodynamics of methane cracking
calculate the difference in the Gibbs free energy, DGC, between
DG0 (Gibbs free energy for methane cracking calculated by
The equilibrium constant for the methane catalytic cracking
reaction is usually expressed only in terms of the partial
pressures of CH4 and H2 [28e30], and assumes that the activity
200 120
of carbon on the catalyst is unity, as shown in Equation (2):
No. of moles per 100 moles CH4 fed

180
    100
KP ¼ p2H2 = pCH4 (2) 160
Methane conversion

eq eq
140
80
However, carbon has been shown to affect equilibrium and 120
its activity seems to be inhomogeneous over nickel. This is due 100 60
to the super-saturation required for carbon filament formation, 80 Hydrogen
40
which further requires different carbon activities at the gas/ 60 Carbon
metal and metal/support interfaces [31e33]. The equilibrium 40 Methane 20
constant for the overall reaction, assuming carbon forming 20 Conversion

a solution in the active metal (e.g. Ni), Ksol


CH4 =H2 , is related to the
0 0
490 540 590 640 690 740 790 840 890 940 990
partial pressure of methane and hydrogen as follows [33]:
Temperature, °C
 
p2H2 csol
CH4 =H2 Fig. 2 e Equilibrium composition and conversion as a
Ksol
CH4 =H2 ¼  eq  (3) function of temperature. The figure shows the equilibrium
pCH4
eq number of moles based on an initial 100 moles of CH4.
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2907

assuming that carbon produced is deposited as graphite, from like in Equation (2) where the carbon activity was assumed
Equations (4) or (5)) and DGObserved (actual Gibbs free energy for equal to 1), a coking threshold, K*M, has been defined and
methane cracking): determined by measuring the reaction quotient, KM ¼
p2H2 =pCH4 , where neither carbon deposition nor gasification
DGC ¼ DGObserved  DG0 (6) occurs, in other words when the overall rate of methane
  cracking is zero [25]. Therefore, when KM > K*M no cracking will
KPobserved take place. Since K*M is measured at zero carbon deposition/
DGC ¼ RT ln (7)
KPGraphite
gasification, it can be assumed that, at this particular condi-
tion, the carbon chemical potential and the solubility of
Rostrup-Nielsen [28] found that the value of DGC decreased
carbon in nickel are uniform throughout the nickel particle. If
with increasing temperature and nickel loading. For 25% Ni/
the carbon deposited was only graphitic carbon, K*M would be
MgO, DGC dropped from 1.88 kcal/mol at 400  C to 0.8 kcal/mol
equal to the KP described above. Snoeck et al. [25] came up
at 600  C. At 500  C, DGC increased from 0.56 kcal/mol for 25%
with mathematical formulas to predict the value of K*M at
Ni to 2.17 kcal/mol for 14% Ni.
different methane partial pressures and temperatures based
De Bokx et al. [35] also studied the thermodynamics of
on experimental observations (see Equations (8)e(12)). For
methane cracking and found a deviation between the calcu-
comparison, they also gave the expressions for the equilib-
lated data and the equilibrium data assuming graphite
rium constants in the case of graphite (Eq. (11)) and nickel
formation for nickel and iron catalysts. Even by taking into
carbide (Eq. (12)).
account the structural disorders and the contribution of the
surface energy of the high surface area filaments [28]; the At PCH4 ¼ 1:5 bar KM ¼ expð116:1=RÞ$expð  100; 765=R=TÞ (8)
deviation between the actual data and the data calculated
based on graphite formation is still considerable [35]. Also, At PCH4 ¼ 5 bar KM ¼ expð156:1=RÞ$expð  134; 230=R=TÞ (9)
Snoeck et al. [25] have differentiated between the enthalpy
and the entropy change calculation based on the experi-
mental results and those calculated using graphite and nickel At PCH4 ¼ 10 bar KM ¼ expð164:7=RÞ$expð  141; 900=R=TÞ (10)
carbide as a final product. They noticed that even the
morphology of the carbon filaments produced from methane gr
Graphite KM ¼ expð104:8=RÞ$expð  84; 400=R=TÞ (11)
cracking is different from graphite and nickel carbide.
Yang and Chen [32] performed solubility measurements of Ni3 C
Nickel carbide KM ¼ expð121:3=RÞ$expð124;600=R=TÞ (12)
carbon in nickel in contact with a mixture of methane and
hydrogen and confirmed that the solubility of carbon in nickel Fig. 3 shows the experimental values for K*M at different
at the support side of the nickel particle was determined by methane partial pressures and temperatures, as well as the
the thermodynamic properties of carbon filaments and that values of K*M for graphite and nickel carbide at different
the solubility of carbon at the gas/metal interface is deter- temperatures. This figure shows a remarkable difference of
mined by the properties of the gas-phase components (i.e. CH4 K*M between graphite and nickel carbide but also shows that
and H2) and the Henry’s law constant of the solution of carbon the values of K*M for graphite and the experimental values at
in nickel. They also reported that the carbon content at satu- 1.5 bar are very close.
ration was 35% higher than that of a mixture in equilibrium Snoeck et al. [25] postulated that for KM < K*M, either
with graphite. encapsulating carbon species or carbon filaments are
Since the chemical potential of carbon is not constant produced. Zhang and Smith [36] determined another
throughout the nickel particle [33], the carbon activity cannot threshold value, KfM, at which only carbon filaments are
be constant and thus cannot be equal to 1. So, instead of using formed. They proposed a criterion for stable cracking activity,
KP to represent the equilibrium conditions for the reaction (e.g. based on the value of KM relative to the two thresholds, K*M

500

1.5 bar
400
5 bar
10 bar
300
Graphite
K*M

Nickel carbide
200

100

0
500 600 700 800 900 1000
Temperatre,˚C

Fig. 3 e KM* versus temperature at different methane partial pressures and KM* for graphite and nickel carbide. Based on
Equations (8)e(12) [25].
2908 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

and KfM. If KfM is between KM and K*M (KM < KfM < K*M) the hydrogen yield and better catalyst stability when copper
cracking activity will be unstable due to the formation of was used as a promoter. Also Figueiredo et al. [26] studied
encapsulating carbon. However, if KM is between KfM and K*M a La2O3-doped NieCu Raney-type catalyst. The catalyst
(KfM < KM < K*M), cracking activity will be stable. showed better stability compared to a La2O3-doped Ni
Raney-type catalyst. Figueiredo et al. [26] attributed the
improved stability to an electronic promotion effect. The
3. Catalysts for catalytic methane cracking improved performance with copper addition was attributed
to the strong influence of copper on the dispersion of the
Most studies considered nickel, iron or activated carbon with nickel by inhibiting the formation of nickel aluminate,
a few papers reporting other active metals, such as cobalt. The which increased the metallic nickel phase available for
support has also been shown to have an effect on the cracking reaction and the nickel surface area subject to reaction. Gac
conversion/deactivation. Therefore, this section will be et al. [40] reported that addition of magnesia to a methane
divided into two parts: active phase and support. cracking catalyst increased the initial cracking rate. The
authors attributed the initial rate increase to the formation
3.1. Active phase of smaller nickel crystals and stronger adsorption sites
when magnesium was present.
This section discusses different catalysts used for methane The catalyst activity is not a linear function of the nickel
catalytic cracking and the different factors affecting catalyst amount in the catalyst/support matrix. Venugopal et al. [41]
activity. The catalyst function is to reduce the activation conducted an experimental study in a fixed-bed reactor
energy required for methane decomposition, leading to lower using a Ni/SiO2 catalyst with nickel loadings in the 5e90%
operating temperatures. Non-catalytic methane cracking is range. The results revealed that increasing nickel loading has
very slow for practical application at temperatures below a positive effect on methane conversion and catalyst stability
1000  C, while catalytic cracking of methane can be conducted until 30% is reached, and there a maximum in conversion was
at temperatures as low as 500  C [19]. For example, for a highly achieved. Increasing the nickel percentage beyond 30%
active catalyst such as Ni/g-Al2O3, optimal performance has resulted in poorer methane conversion and catalyst stability,
been reported in the 500e550  C temperature range [37]. as shown in Fig. 4.
Iron group metals are known to have the highest activity Due to fast deactivation observed with nickel-based cata-
for hydrocarbon cracking. In the case of methane, which has lysts at higher temperature (w600  C) [42], different catalysts
the highest stability in comparison to other hydrocarbons [9], have been tested either as a mixture with nickel, like copper
nickel has been described as the most active catalyst for and iron [20e42], or separately, for example iron [43], copper
methane cracking among the iron group metals. Cobalt could [44], and activated carbon [2,27,45]. Chesnokov and Chichkan
potentially be used as a catalyst for methane cracking, but it is [42] developed a 70%Nie10%Cue10%Fe/Al2O3 catalyst for
less active, has toxicity issues, and higher cost compared to methane cracking, where the addition of iron increased the
nickel. Iron has also been studied as a catalyst for methane optimal operating temperature range from 600 to 675  C for
catalytic cracking but it showed lower activity compared to Ni/Cu/Al2O3 to 700e750  C while maintaining good catalyst
nickel [19]. Direct comparison shows that the methane stability. Jang and Cha [43] used an iron supported on alumina
cracking catalytic activity for the iron group metals is: catalyst. The performance of the iron catalyst was greatly
Ni > Co > Fe [38]. affected by the reaction temperature; at temperatures below
Operating conditions, such as temperature, electronic state 750  C, the catalyst rapidly deactivated, but at higher
and dispersion, affect nickel’s stability and activity for temperatures, the catalyst retained a reasonably stable
methane cracking [20,39]. Optimum operation has been activity for at least 100 min. At 1000  C, Jang and Cha [43],
observed in the 500e552  C temperature range [19], and in
a separate study, a maximum in methane conversion was
observed at 552  C [35]. The maximum conversion being
attained at 552  C is due to equilibrium being achieved
between the carbon deposited on the nickel and the surface
migration and diffusion of carbon through nickel. At
temperatures higher than 552  C, carbon is formed at a rate
higher than bulk diffusion and surface migration causing an
excess of carbon to deposit as encapsulating carbon [19,37]. At
temperatures below 500  C, carbon is formed at a rate lower
than the carbon solubility in nickel, thus reducing the driving
force for carbon diffusion in the metal. The consequence is
that less carbon is formed before the catalyst deactivates [37].
Nickel-based catalyst performance is not just a function
of reactor operating conditions; it is also a function of nickel
electronic state and dispersion. For example, Echegoyen
et al. [20] studied the role of copper on nickel dispersion by Fig. 4 e Methane conversion over various Ni/SiO2 catalysts
adding 3% copper nitrate to a nickel/alumina catalyst during at 600  C (reprinted from [41], with permission from the
catalyst preparation; the results showed an enhanced International Journal of Hydrogen Energy).
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2909

using a fluidized bed, reported an increase in activity over methane conversion achieved using silica is attributed to the
time with an iron catalyst; over 100 min the methane instability of nickel silicates, which may form, and their
conversion continuously increased, reaching 90%, but no data tendency to decompose at higher temperatures during the
were presented past 100 min. Ammendola et al. [44] have used reduction step. In another study, Ermakova and Ermakov [9]
a copper supported on alumina catalyst for methane cracking compared Ni/SiO2 and Fe/SiO2 using different catalyst prepa-
and were able to achieve, in a fluidized bed, 90% methane ration methods and operating conditions. For nickel, the
conversion for 90 min, but for very diluted methane condi- maximum yield (384 g C/g Ni) was achieved when the catalyst
tions (i.e. 5%). Iron oxide has been studied as an alternative, was silicate free; if 1.5e2% of the nickel was transformed into
but the results showed no appreciable conversion below nickel silicate the yield dropped to 40 g C/g Ni. For Fe, the
800  C [2]. addition of silica resulted in inhibition or promotion of
Activated carbon has also been tested as a catalyst for methane conversion depending on the amount of silicate in
methane cracking in fluidized and fixed-bed reactors. In both the catalyst formed during the pre-treatment. Takenaka et al.
reactors, activated carbon showed similar behaviour, with an [47] used X-ray diffraction to characterize nickel on different
initially high catalytic activity, but then rapid deactivation due supports (SiO2, TiO2, graphite, Al2O3, MgO and SiO2$MgO). All
to carbon deposition. The deposited carbon was shown to catalysts were calcined in air at 600  C for 5 h, and reduced
block access to active sites [2,27]. Suelves et al. [45] used with hydrogen at 550  C for 1 h. They found that, for equiva-
carbon black and activated carbon as catalysts for methane lent surface area, the lower the interaction between nickel
catalytic cracking and reported better performance with and the support, the higher the methane conversion. Finally,
carbon black, while the activated carbon suffered from rapid under their test conditions, they showed that silica and titania
deactivation. The authors found that pore distribution and the supports resulted in the highest methane conversions, as
surface chemistry played an important role in the instanta- shown in Fig. 5.
neous rate and catalyst lifetime before deactivation. The support structure and textural properties (e.g.
The catalyst pre-treatment method is another important porosity) also affect methane conversion. Ermakova et al. [19]
factor for performance, primarily affecting the percentage of studied the catalytic performance of nickel promoted with
the metallic catalyst that will be active during the reaction. different supports, including silica, magnesia, alumina, and
After the catalyst is prepared using impregnation, co-precip- zirconium oxide. The results indicated that the catalytic
itation or any other preparation techniques, the catalyst needs activity and lifetime of the catalyst depend significantly on the
to be activated (nickel metal is the oxidation state required) pore structure of the catalyst; they reported that nickel
prior to reaction. The pre-treatment includes calcination and promoted with silica (i.e. with silica as a promoter, not
reduction. The calcination step helps in removing any residual support) that has wide pores gave the highest methane
precursor species and to change the precursor material to conversion and longest lifetime.
metal oxide. The reduction step is required after calcination to The support structure may also affect the outlet gas
transform metal oxide to the active metal phase. The composition and the morphology of the deposited carbon. For
temperature and the duration of the calcination and reduction example, higher oxygen capacity supports like ceria, are not
steps influence how much catalyst is activated. Calcination or favorable unless the catalyst is prepared in such a way that
reduction above the optimum temperature also affects the the surface oxygen is immobilized to prevent it from reacting
catalyst texture and may cause sintering, whereas using with deposited carbon, and forming carbon monoxide [48].
temperatures lower than the optimum temperature may not
activate the catalyst completely. For methane catalytic
cracking, Echegoyen et al. [20] studied the effect of calcination
temperature on nickel activity during reaction and found that
600  C resulted in the highest yield for the catalyst tested. Li
et al. [46] reported that nickel supported on g-alumina was
reduced easily between 500 and 920  C, but the lower the
reduction temperature, the better the ultimate catalyst
activity [10].

3.2. Support material

Methane conversion is a function of the catalyst matrix, i.e.


including the active material and the support [40,41]. Eche-
goyen et al. [20] found that minimal interaction between the
active component and the support is important for increased
conversion. The support material also directly affects
methane conversion by affecting the surface area of metal
subjected to the reaction and the metal’s electronic state. For
example, the effects of magnesia and silica as supports for
nickel were compared [19]. The authors attributed the lower Fig. 5 e Methane conversion over Ni catalysts supported on
methane conversion using magnesia to the formation of different supports at 500  C (reprinted from [47], with
a solid solution between Ni and magnesium, while the higher permission from Elsevier).
2910 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

4.2. Nucleation
4. Description of carbon filament formation
and growth Since most studies were conducted with nickel as the active
metal, the nucleation discussion presented will be primarily
4.1. Overview of the process for carbon formation relevant to nickel, although most of the processes describe
here could be also applied to other metals.
There are three types of coke that form during methane The first stage of filament growth is nucleation. During this
cracking on supported metal catalysts, namely polymeric, stage nickel possesses liquid properties and the nucleation is
filamentous, and graphitic carbon. Polymeric coke is produced due to 1) high carbon coverage on the Ni surface, 2) subse-
during thermal decomposition of hydrocarbons, while the quent high concentration of carbon dissolved in the nickel
filamentous and graphitic carbons are formed during catalytic particle and 3) carbon segregation behaviour.
cracking of hydrocarbons [34,49]; the classification here is The surface coverage of carbon at the gas side is increased
related to the reaction conditions [49]. Coke can also be to high levels and, as a result, the net rate of surface reactions
characterized based on its reactivity with hydrogen, water, is reduced, even though the gas mixture may have a compo-
and oxygen [49]. The formation of coke on nickel is a multi- sition that thermodynamically favours carbon formation.
stage process. The process starts with methane adsorption Then, the carbon dissolves in the nickel particle, diffuses
followed by several dehydrogenation reactions and ends with through it, and precipitates at the metal/support interface
coke formation, as shown in Fig. 6. After methane is adsorbed [51,52]. According to Kock et al. [52], a key step prior to
on the fresh nickel catalyst and the hydrogen is released, nucleation is the formation of nickel carbide. They proposed
the remaining deposited carbon dissolves in nickel forming the following nucleation mechanism in the case of methane
a layer of uniform concentration solution of carbon in nickel cracking over a 50 wt% Ni/SiO2 catalyst:
at the gas side; the carbon concentration is equivalent to the
gas-phase solubility of carbon in nickel. Then carbon diffuses 1. Carburization to form nickel carbide
through the nickel particle to the support side until the carbon 2. Carbide decomposition
solution of nickel is super-saturated with respect to the fila- 3. Nucleation of graphite
mentous carbon; at that time, nucleation of filamentous 4. Precipitation of graphite layer at the nickel/support
carbon begins [32,33,50,51]. With carbon filament growth, interface
encapsulating carbon forms, which slows the cracking rate 5. Detachment of the nickel particle from the support.
and ultimately deactivates the catalyst.
The driving force for filament formation is the tempera- The detachment is responsible for the appearance of the
ture gradient [32,51] and/or the concentration gradient due carbon filament. Solid or hollow filaments of various diameters
to the difference in solubility of carbon at the gas/metal are formed depending on operating conditions, in particular
interface and that at the metal/support interface [29,31,52]. the temperature [33]. At low temperatures, diffusion through
It can be seen from Fig. 6 that at the gas/metal interface, the metal particle is slower than the nucleation/precipitation
carbon is in the form of a solute dissolved in nickel as rates and thus nucleation occurs more uniformly at the metal/
a solvent, forming an unstable intermediate of surface support interface. As a consequence, the detachment step
nickel carbide or “selvedge”, which is super-saturated with leads to a non-hollow solid filament. In addition the nickel
carbon at high temperature and the carbon concentration particle retains most of its original shape. On the other hand, at
and temperature are decreasing in the direction of metal/ higher temperatures, distortion of the nickel particle has been
support interface. observed [33,53,54] when the nucleation rate becomes greater
than the diffusion rate. In this case, nucleation/precipitation
occurs close to the metal/gas interface, thus creating a hollow
filament. The hollow feature is due to the uplifting of the
particle where no precipitation occurred. At these particular
locations, because nickel/support binding forces must be
overcome, a deformation of the particle occurs forming
a “pear”-shaped particle [33]. Schematics of the detachment
with formation of solid and hollow carbon filaments, along
with particle deformation, are shown in Fig. 7a and b, respec-
tively [33]. Other shapes have also been proposed, such as
a conical shape [53]. Boellaard et al. [53] investigated the
microstructure of the carbon filaments and mentioned that
the coke layers resemble fish-bone-like structures, which were
also discussed by Gac et al. [40], and proposed a filament
growth mechanism where carbon species are excreted from
the metal particle perpendicular to the metal/filament inter-
face and deposited upon one another in a conical form, as
Fig. 6 e Schematic of the classical mechanism of carbon shown in Fig. 8a. However, this mechanism assumed that the
filament formation (reprinted from [25], with permission perimeter of the cone increases and eventually leads to the
from Elsevier). filament bursting, which is not consistent with the assumption
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2911

catalyst, the filament diameter decreases with increasing


temperature [24], as shown in Fig. 9. It was postulated that the
change in nucleation rate and particle deformation at various
temperatures may be responsible for the differences in
diameter observed.
Catalyst characteristics (i.e. preparation method, metal
loading, and calcination and reduction temperatures), as well
as gas composition control the maximum number of fila-
ments that can be formed and consequently the maximum
amount of carbon that can be formed during the nucleation
step [13]. For example, increasing the methane concentration
accelerates the nucleation rate and the number of filaments
formed since the partial pressure of methane determines the
degree of super-saturation required for nucleation to begin
[13,33]. The degree of super-saturation required for nucleation
is exceeded only when the activity of carbon in the gas phase
exceeds unity and the activity of carbon, aC, can be calculated
using the following equation [32]:

p2H2
aC ¼ KP  (13)
pCH4

Experimental observations have shown that the nucleation


of filamentous carbon is more difficult when there is a lower
partial pressure of methane. This is due to slower nucleation
and longer periods over which the rate of carbon formation
increases, resulting in a smaller number of carbon filaments
formed [32].
Fig. 7 e Particle detachment and formation of a) solid Also, Villacampa et al. [13] showed that as the catalyst
filament and b) hollow filament. Note the particle reduction temperature increases more nickel sintering occurs
deformation in b (reprinted from [33], with permission and fewer filaments are formed. Yet, because of the lower
from Elsevier). interaction between the nickel and support at higher reduc-
tion temperature, the nucleation rate still increases with the
reduction temperature, which indicates the presence of an
optimal reduction temperature.
that the diameter of the filament remains constant over very
long distances. In order to overcome this problem they added
the concept of slippage in another mechanism they proposed, 4.3. Filament growth
where the excreted coke layers push metal particles upward
along the direction of the filament axis, so that the filament After the nucleation period, the Ni particle detaches from the
diameter remains constant as shown in Fig. 8b. support surface, now being supported by the carbon filament,
Temperature also has an effect on the filament diameter, and the concentration of the carbon at the metal/filament
where for temperatures between 500 and 650  C for a Ni/Al2O3 interface drops to the saturation concentration of filamentous

Fig. 8 e Cross sections of conical graphite layers excreted in a direction perpendicular to the metal/filament interface;
a) without slippage, b) with slippage (reprinted from [53], with permission from Elsevier).
2912 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

Fig. 9 e SEM micrographs of a 5 wt% Ni/a-Al2O3 catalyst after (a) 14 h reaction at 650  C, (b) 13 h reaction at 600  C, (c) 8 h
reaction at 550  C and (d) 16 h reaction at 500  C (reprinted from [24], with permission from Springer).

carbon, leading to a different carbon concentration gradient in turn, slows down the removal of carbon at the gas side which
the nickel particle, and thus a different rate of carbon diffu- further increases the rate of encapsulating carbon deposition
sion [33]. A thermal phenomenon has been reported imme- [51]. Because of these cause-and-effect phenomena, the active
diately following nucleation: after detachment, the nickel is nickel particles ultimately become completely encapsulated
on the tip of the insulating carbon filament and the support no by carbon and filament growth ceases.
longer acts as a heat sink [40,51]. As a consequence, the
particle temperature during filament growth increases
4.5. Proposed mechanisms for carbon filament growth
somewhat, thus also contributing to an increase in carbon
growth rate [51].
Although most authors agree on the main stages for carbon
A constant filament growth rate is reached when
filament growth, different mechanisms have been proposed
the maximum number of filaments is nucleated. During the
for carbon filament formation, with different assumptions
constant growth rate stage, a uniform temperature on the
concerning the composition and nature of the selvedge and
nickel particle prevails due to the balance between the heat
the driving force for carbon movement. In the following
losses and the heat gains from the surrounding environment,
paragraphs, the different mechanisms of carbon filament
and the number of filaments remains constant [55].
formation are discussed in their chronological order of
appearance in the literature.
4.4. Catalyst deactivation or tailing stage Assuming temperature as the driving force for carbon
filament formation, Baker et al. [51] developed a three-step
Due to the presence of excess carbon on the gas side of model to explain filament formation. In the first step, hydro-
the nickel surface, some carbon deposits on the nickel carbons adsorb and decompose on certain faces of the metal
surface form encapsulating carbon, which is responsible for particle. The second step involves dissolution of some of the
catalyst deactivation. As the encapsulating carbon deposits on carbon species into the bulk and diffusion through the metal
nickel, a gradual decrease in available nickel surface area for particle from the hotter leading face (exposed to the gas) to the
methane cracking occurs. In addition, this gradual decrease in cooler rear face (facing the support), where carbon is precipi-
active surface area results in less heat input to the metal tated from the solution to form carbon filaments. Finally, in
particle, which decreases the rate of carbon diffusion through the third stage, the growth rate declines due to the formation
the nickel particles by reducing the carbon solubility. This, in of encapsulating carbon.
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2913

McCarty et al. [50] assumed that an intermediate product, catalyst was gradually encapsulated by the free carbon
called selvedge, is formed at the gas/metal interface due to the thereby isolating the catalyst from methane.
segregation behaviour of carbon in nickel. The segregation Toebes et al. [59] studied carbon deposition by decomposi-
behaviour takes place due to the carbon concentration tion of carbon-containing gases (CH4, CO/H2 or C2H4/H2) using
gradient between the surface and bulk over a number of Ni/SiO2 or unsupported nickel as a catalyst. They assumed
atomic layers. Then, filamentous carbon is formed as a product a delicate balance between dissociation of the carbon-con-
of hydrocarbon dissociative chemisorption on the metal taining gases and carbon diffusion through the catalyst
surface. The dissociative chemisorption produces carbon particle, and a balance between carbon diffusion through the
atoms on the catalyst particle face, which then diffuse towards catalyst particle and rate of nucleation and formation of
the opposite face where the carbon atoms crystallize in the graphitic layers. As in most of the other studies, the authors
form of a continuous graphite-like structure. Based on their proposed that the carbon filament is formed by carbon diffu-
evidence they concluded that this rate of carbon deposition is sion through nickel from the gas side to the support side. The
governed by isothermal carbon diffusion through the metal diffusing carbon is deposited at the nickel/support interface
particle. and leads to carbon filament formation. The authors support
Kock et al. [52] discussed carbon thermodynamic proper- the mechanism that describes the segregation behaviour of
ties as a function of carbon origin. They proposed that the carbon in nickel, creating a selvedge with a high concentration
carbon at the metal/filament interface has a higher solubility at the nickel surface on the gas side. They again suggested
in the catalyst than the carbon deposited at the metal/support there is a gradient in carbon concentration with a high level at
interface; and the difference in solubility is the driving force the surface and less in the bulk. Furthermore, the segregated
for carbon transport. carbon atoms at the surface compete with the gas-phase
Alstrup [29] defined the selvedge as surface carbide. This atoms (methane and hydrogen) for active sites. They also
assumption of carbide presence was also used by Schouten indicated, as mentioned above with other studies, the carbon
et al. [56] who assumed that the carbon atoms entering the diffusion driving force is the difference between dissolved
selvedge create surface carbide species. But Snoeck et al. [33] carbon in the nickel at the gas side of the particle and solubility
reported that assuming nickel carbide as an intermediate of carbon at the support side of the particle.
product is impractical because nickel carbide decomposes at In terms of carbon species activities, not isolated to fila-
350  C and the presence of sub-stoichiometric carbide could ments in this case, different types of carbon are produced
not be determined with certainty during steady-state carbon during hydrocarbon dissociation, each with its own properties
filament formation. and reactivity [34]. Trimm [60] mentioned in his review of
Instead, Snoeck et al. [33] developed a model for carbon coking during steam reforming that two kinds of carbon are
filament formation that was based on the solubility of carbon produced when hydrocarbons dissociate, namely an a-type
in nickel depending on the extent of the driving force in the carbon (Ca) and a b-type carbon (Cb). The Ca is more reactive
gas phase towards carbon formation (i.e. higher partial pres- than Cb and can be easily gasified using oxygen. When an
sure of methane and/or lower partial pressure of hydrogen). excess amount of Ca is present, Ca transforms to Cb. Although
The higher the driving force towards carbon formation, the the fate of Cb has not been completely elucidated, it is thought
higher the amount of dissolved carbon in nickel. Snoeck et al. that some of the Cb may not diffuse through the nickel but
[33] found that at a fixed temperature in the nickel particle, the instead, remain on the catalyst surface, slowly building up
solubility at the gas/metal interface and solubility at metal/ and eventually encapsulating the catalyst leading to complete
filament interface may differ by a factor of 200. deactivation. Guo et al. [49] used temperature-programmed
Aiello et al. [57] used a mechanism similar to the previously reaction techniques and Raman spectroscopy to characterize
mentioned ones, but assumed that the rate-determining step carbon deposited on 5% Ni/MgAl2O4 during methane cracking.
is carbon diffusion through nickel and that catalyst deacti- They found three forms of carbon on the catalyst surface
vation results from the space limitation for carbon filament during methane catalytic cracking, a-type carbon (Ca) and
growth. They assumed that deactivation takes place when the b-type carbon (Cb), and g-type carbon (Cg). Using Raman
carbon filaments come into contact with each other. spectroscopy, they found that Cg is graphite-like, and assigned
Shah et al. [58] studied the catalytic decomposition of pure it as that responsible for catalyst deactivation. During carbon
methane using nano-scale binary FeeM (M ¼ Pd, Mo, or Ni) gasification experiments, they reported that Cg showed
catalysts supported on alumina. SEM and TEM techniques a resistance to gasification using oxygen or hydrogen. However,
were used to characterize carbon. A substantial amount Cg showed an unexpectedly high gasification rate with carbon
(>90%) of the carbon produced during catalytic decomposition dioxide, which they attributed to the formation of carbonate,
of methane, between 700  C and 800  C, was in the form of bidentate and formate species on MgAl2O4. Other authors (e.g.
potentially useful multi-walled carbon nanotubes, but the [11,34]) identified five different types of carbon that may form
carbon formed above 900  C was amorphous with a small during hydrocarbon dissociation, which can be distinguished
percentage of carbon flakes and carbon fibres [40]. Shah et al. based on their crystalline properties and reactivity, and are
[58] confirmed with the help of XAFS and Mössbauer spec- listed as:
troscopy that the catalyst particle remained in its metallic
state, even after it had been lifted off the alumina support, and 1. Ca: Adsorbed atomic carbon (dispersed, surface carbide),
was covered by encapsulating carbon. No carbide species were formed between 200 and 400  C.
observed. Shah et al. [58] noted that the change in surface 2. Cb: Polymeric filaments (amorphous), formed between 250
structure did not actually deactivate the catalyst; instead the and 500  C.
2914 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

3. Cv: Carbon filaments (amorphous), formed between 300 Some researchers attribute deactivation to reasons other
and 700  C. than the carbon layer formed around the active sites. For
4. Cg: Nickel carbide (bulk), formed between 150 and 250  C. example, Ishihara et al. [64] observed that a 10% Ni/SiO2
5. Cc: Graphitic film (crystal), formed at temperatures greater catalyst was still active after a large amount of carbon
than 600  C. deposited on the surface, up to a 200-carbon atom/nickel atom
ratio. They concluded that active site blocking is not neces-
sarily the reason for deactivation. Using scanning electron
5. Catalyst deactivation microscope (SEM) and transmission electron microscopy
(TEM) to characterize the spent catalyst, they found that
5.1. Coking as the main deactivation process in methane carbon was deposited as filaments with the nickel particle on
catalytic cracking the tip, as described above. Zhang and Amiridis [37] concluded
that deactivation occurred due to space limitation, when the
Ni-based catalysts are highly active and selective for hydro- formed filaments began to interfere with each other inside or
carbon cracking. They are characterized by a high capacity for outside the catalyst pellet, inhibiting the deposition of more
carbon adsorption [49], but the adsorbed carbon also causes carbon atoms in filamentous form. Aiello et al. [57] reported
deactivation. There have been multiple studies focused on the the same conclusion when they studied methane cracking
deactivation mechanisms with the ultimate goal targeting using Ni/SiO2; after deposition of thousands of carbon atoms
lengthening the catalyst life. These studies have evaluated on nickel, the catalyst was deactivated due to space limita-
parameters affecting deactivation, the period of stable cata- tions imposed by reactor space.
lyst performance, and when the catalyst needs to be replaced.
Poisoning, fouling or coking (carbon deposition), sintering and 5.2. Coking deactivation mechanism
mechanical degradation are the common forms of catalyst
deactivation [11]. Methane decomposes at the gas/metal interface, producing
The main deactivation mechanism during methane cata- carbon which dissolves and diffuses through and around the
lytic cracking is coking, which can be defined as the physical catalyst particle to the metal/support interface, detaching the
deposition of carbonaceous species from the reacting species catalyst particle from the support, forming a filament with
onto the catalytic surface [11,61,62], which may result in loss the catalyst particle on the filament’s tip [39,40,57,64].
of catalytic activity as a result of blocking of catalyst sites and/ Diffusing carbon deposits as filamentous carbon [39], and the
or pores [11,63]. When large amounts of carbon deposit, deposition rate is controlled by heat transfer through carbon
coking may also result in disintegration of catalyst particles [9] or the concentration gradient of carbon through the nickel
and reactor plugging [11]. The coke/carbon formed during particle [29]. Carbon diffusion through the metal particle has
hydrocarbon processing is usually classified according to been proposed as the rate-controlling step, and deactivation is
reaction type, catalyst type, and reaction conditions [63]. a result of space limitation in the reactor [57,64] and/or is
Deactivation due to coke forming reactions on metals (e.g. a result of a solubility limitation of carbon in nickel when the
methanation, FischereTropsch synthesis, and steam reform- dissolved carbon reaches its saturation concentration, which
ing) occurs due to the difference between the carbon forma- leads to encapsulating carbon formation [24,33,40,58]. As
tion and carbon gasification rates [63]. Coking may affect a result of carbon deposition on nickel, the catalyst activity
catalyst activity in several ways; in some cases all the effects changes as the reaction proceeds.
combine [11,34]. The carbon can: So, during coking-based deactivation, carbon is deposited
on the active site at the active site/gas interface forming a layer
1. adsorb strongly on the active phase surrounding and called encapsulating carbon, which blocks reactant access [36].
blocking access to the active phase surface; Encapsulating carbon causes catalyst deactivation either by
2. encapsulate the active metal particle; blocking the catalyst pores or by complete encapsulation.
3. plug the micro and mesopores, denying access to the active Encapsulating carbon formation has been found to be
phase inside the pores; inversely proportional to methane partial pressure, while
4. accumulate as strong carbon filaments leading to catalyst increasing hydrogen partial pressure has a negative impact
pellet disintegration; and on the encapsulating carbon formation rate [13,26]. Also,
5. in extreme cases, physically block the reactor. increasing temperature increased encapsulating carbon
formation [44]. The form of the encapsulating carbon is also
The deposited carbonaceous materials may vary in chem- dependent on the temperature. At temperatures <500  C,
ical structure from hydrogen-deficient, aromatic-type poly- encapsulating carbon is deposited as a polymeric film. While at
mers, to graphitic carbon. In extreme cases the catalyst temperatures >450  C, deposition of whisker-like carbon is
surface is covered with layers of coke deposit, thereby more likely to take place. But at higher temperatures >600  C,
decreasing the accessibility to surface sites. Although coking pyrolitic coke forms as the encapsulating carbon layer [34,63].
is not desirable in most hydrocarbon processing reactions, in Takenaka et al. [47] studied the structural changes of nickel
some catalytic cracking processes, for example the catalytic during methane cracking over Ni/SiO2. The XANES/EXAFS
cracking of methane for hydrogen production, coking is spectra of the pure nickel were measured, and compared to
a direct product. More coke deposited on the catalyst means the spectra of the supported nickel to measure the changes in
better hydrogen production occurred, since the reaction nickel structure during methane cracking. XRD was also used
products are hydrogen and coke. to determine the chemical forms of nickel. The results showed
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2915

that nickel kept its metallic state until the ratio of the carbon deactivation rate is inversely proportional to hydrogen partial
deposited over nickel was, C/Ni, >900, as shown in Fig. 10. The pressure. We can summarize the effect of different factors on
results also indicated that the methane cracking rate was not the reaction rate and deactivation rate as follows: any
affected as long as nickel remained in its metallic form. After parameter that can increase the reaction rate also increases
the C/Ni ratio exceeded 900, XANES/EXAFS spectra showed the deactivation rate. Ermakova et al. [19] studied the effect
some discrepancy between the catalyst and nickel foil read- of temperature on deactivation. Their conclusions agreed
ings, whereas the catalyst spectrum before reaction resem- with that of Suelves et al. [65]. An example of the effect of
bled that of Ni foil. In addition, the methane cracking rate temperature [19] on deactivation rate is shown in Fig. 11.
started to decrease for C/Ni ratios above 900. The XANES Suelves et al. [65] studied the methane flow rate and
spectrum for the supported nickel was then closer to that for temperature effects on deactivation using a fixed-bed reactor
nickel treated with CO, and indicated complete deactivation. and a 65% Ni supported on silica and alumina catalyst, with 2
XRD studies of the deactivated catalyst showed that nickel and 0.3 g of catalyst used to vary the space velocity. Using
was present in two forms; metallic and carbide. The authors 100% methane and at 700  C, Suelves et al. [65] found that with
concluded that the carbide form is responsible for catalyst a space time of 1.15 s, catalytic activity did not decay after 16 h
deactivation via formation of an inactive component with on stream, with a carbon yield of 55.6% (as a percent of the
nickel. The authors also studied the detected carbide form, total methane passed through the reactor) and 6.3 g of carbon/
and the results indicated that the physical properties of the g of catalyst. While at a space time of 0.23 s, the catalyst
detected carbide are different from that of Ni3C. completely deactivated after 90 min with almost half
the amount of carbon deposited, about 6.7% yield. Then the
5.3. Factors affecting catalyst deactivation authors used the 1.15 space time but at 550  C to study the
effect of temperature. The results showed that at 550  C,
The catalyst deactivation rate is of course affected by oper- the catalyst held its activity for 16 h at a yield of 43.8% and
ating conditions. The operating conditions reported in the a total amount of 14.3 g of carbon/g of catalyst. In their
literature as having the most effect on the deactivation rate conclusions, they reported that increasing the flow rate and/
are methane flow rate (or more specifically the gas hourly or the reaction temperature decreased catalyst lifetime.
space velocity, GHSV), reaction temperature [19,65], methane The third factor affecting catalyst deactivation is the
partial pressure, and hydrogen partial pressure [13]. Suelves hydrogen partial pressure, which was studied by Villacampa
et al. [65] reported that the higher the temperature and et al. [13]. They used different mixtures of CH4 and H2 to study
methane flow rate, the shorter the catalyst life; so, tempera- the influence of hydrogen on the kinetics of methane catalytic
ture and methane flow rate are directly related to the catalyst cracking. They observed that the presence of a small amount
deactivation rate. Villacampa et al. [13] found that the of hydrogen actually increased the total amount of carbon
deposited, although the methane cracking rate was lower,
compared to the hydrogen-free feed, as shown in Fig. 12.
Toebes et al. [59] attributed the effect of hydrogen partial
pressure to a reduction in the encapsulating carbon formation
rate. Because hydrogen is a reaction product, and may adsorb
on, or interact with, the active sites, it can prevent methane
from adsorbing, which reduces both the cracking rate and the
deactivation rate.
Villacampa et al. [13] studied the effect of methane partial
pressure on catalyst deactivation and carbon deposition rate.
Increasing PCH4 increased cracking and deactivation rates, and
slightly reduced the total carbon deposited on the catalyst.

Fig. 10 e Ni K-edge XANES/EXAFS spectra of Ni/SiO2


catalysts with and without deposited carbon, Ni foil,
and the foil treated with CO (reprinted from [14], with Fig. 11 e Methane conversion vs. time, at 600  C (squares),
permission from the International Union of 550  C (circles), and 500  C (triangles) (reprinted from [19],
Crystallography (http://journals.iucr.org/)). with permission from Elsevier).
2916 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

is better (e.g. air regeneration). On the other hand, additional


hydrogen is produced during steam regeneration, approxi-
mately 2 and 3.4 moles of hydrogen per mole of methane that
was introduced during the cracking portion of the cycle for air
oxidation and steam gasification, respectively [37,57]; but
steam regeneration is endothermic while air regeneration is
exothermic, as shown in Equations (14) and (15) [5]:

C þ O2 4CO2 ; DH1073 ¼ 174:5 kJ=mol (14)

C þ H2 O4CO þ H2 ; DH1073 ¼ 135:9 kJ=mol (15)

Regarding the regeneration time, the faster regeneration


process is obviously better so higher catalyst circulation rates
can be achieved and the air regeneration process is faster than
steam regeneration. Regarding the effect on catalyst texture
however, in air regeneration, local hot spots may form,
affecting some active sites by converting them to an oxide
form, which would then require a reduction step before using
the catalyst again for methane cracking [37]. Also, and prob-
ably more serious, the localized high temperature areas may
cause sintering, decreasing the exposed nickel surface area
[57]. On the other hand, in steam regeneration, the catalyst
bed temperature can be kept more uniform, thereby avoiding
sintering [37,57]. Furthermore, Muradov et al. [66] mentioned
that treatment with steam may result in an increase in cata-
lyst surface area for activated carbon. Muradov et al. [66] used
activated carbon as a catalyst for methane cracking at 850  C,
Fig. 12 e Influence of hydrogen partial pressure on (a) and after 1 h the catalyst partially deactivated and the
carbon formation rate and (b) carbon content deposited surface area of the activated carbon dropped from 670 m2/g to
on the catalyst (reprinted from [13], with permission from 324 m2/g. After treatment of the activated carbon with steam
Elsevier). at 950  C for 0.5 h, the surface area increased to 617 m2/g.

6.1. Air regeneration


The deactivation rate is higher with increasing PCH4 due to the
Zhang and Amiridis [37] used a 16.4% Ni/SiO2 catalyst for
increase in encapsulating carbon formation.
methane cracking and air oxidation to regenerate the spent
catalyst at 550  C. The catalyst was quickly and fully regen-
erated using air oxidation but the high temperature generated
6. Catalyst regeneration during the oxidation process resulted in disintegration of the
sample to a fine powder. This was also reported by Rahman
The major challenge for a continuous hydrogen production [67]. Furthermore, XRD analysis showed that Ni oxide had
process using catalytic cracking of methane is regeneration of formed, requiring reduction of the catalyst before further use
the deactivated catalyst. Two different catalyst regeneration for methane cracking.
methods are commonly proposed in the literature: steam Villacampa et al. [13] used a 3% oxygen stream to fully
regeneration and air regeneration, and to a lesser extent CO2. regenerate a 30% Ni/Al2O3 catalyst. They reported that the
In steam regeneration, steam reacts with carbon producing catalyst lost its activity after the first regeneration experi-
hydrogen with a mixture of carbon monoxide and carbon ment, as shown in Fig. 13. Rahman et al. [24] found very
dioxide. In air regeneration, oxygen reacts with the deposited similar results. Villacampa et al. [13] attributed the loss of
carbon to burn it; a mixture of carbon oxides is produced activity after regeneration with air to active site sintering due
depending of the amount of excess air used. In carbon dioxide to the high temperature generated during oxidation of the
regeneration, deposited carbon reacts with CO2 forming CO, deposited carbon. Rahman [67], on the other hand, attributed
producing a mixture of carbon oxides. Each method has the activity loss to the fact that the catalyst was completely
its advantages and disadvantages, and the choice of the disintegrated after the first regeneration, as seen in Fig. 14,
optimum method depends on the overall process economics. where it is apparent that the catalyst became like fine powder.
Three factors are essential before choosing the regeneration It was speculated that this may be due to the growth of carbon
method; the energy needs for the process, the catalyst circu- filaments inside the porous catalyst during the previous
lation time, and the effect on catalyst performance. cracking portion of the cycle.
From an energy point of view, a method that will produce Otsuka et al. [5] used oxygen to regenerate the carbon
energy to cover part of the methane cracking energy demand deposited on PdeNi/SiO2, Ni/TiO2, and Ni/Al2O3, and reported
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2917

As an alternative to complete regeneration, the concept


of partial regeneration was introduced [24,68]. In partial
regeneration, only part of the deposited carbon is removed.
Koc et al. [68] partially regenerated a 15%Ni/g-Al2O3 catalyst at
500  C using air and found that the partially regenerated
catalyst, after gasifying 90% of the coke, regained excellent
activity, similar to that during the previous cycle, for up to
4 cycles.
Besides sintering and disintegration, deterioration in
performance after complete regeneration was also attributed
to the change in orientation of nickel atoms from Ni (110) to Ni
(111) and Ni (100) [65]. Ni (110) is a highly active face for carbon
formation while Ni (100) has much lower activity, and Ni (111)
shows no activity at all [69].

6.2. Steam regeneration

For steam regeneration of the spent catalyst, the process can


be described as steam reforming with two stages. In the first
stage only pure hydrogen is produced, while in the second
stage hydrogen contaminated with carbon monoxide and
carbon dioxide is produced [37].
Zhang and Amiridis [37] used steam gasification to regen-
erate a deactivated 16.4% Ni/SiO2 catalyst at 550  C. The cata-
lyst regained its activity and the XRD pattern showed that the
nickel kept its metallic form after regeneration was complete,
but also indicated the presence of small pockets of carbon.
Fig. 13 e Influence of regeneration cycles on the evolution TEM was used to investigate the nature of the remaining
of (a) hydrogen production rate and (b) on carbon content carbon. The data indicated that the remaining carbon was
using 30% Ni/Al2O3 and the cracking temperature was filamentous, but had thinner walls and the authors concluded
600  C (reprinted from [13], with permission from from this that the external walls of the originally solid fila-
Elsevier). ments were more resistant to steam gasification. Zhang and
Amiridis [37] suggested running an air oxidation cycle with
successive cycles of steam gasification to remove the steam
gasification resistant carbon.
that all the used catalysts regained their activity after regen- Aiello et al. [57] regenerated a deactivated 15% Ni/SiO2 at
eration, for 5 cracking/regeneration cycles. Carbon deposited 650  C using steam, and regained initial activity. The results
on the catalyst was oxidized at 480  C for those tests, but for revealed no appreciable loss in catalytic activity after 10
complete elimination of the deposited carbon, a temperature successive cycles, as shown in Fig. 15. XRD analysis indicated
higher than 500  C was needed. They also ran catalyst activity no increase in carbon remaining on the catalyst after
tests at 550  C for all the used catalysts for several cracking/ successive regenerations and no structural changes in the
regeneration cycles. For Ni/SiO2, the catalyst lost its activity nickel particles as the catalyst was cycled between cracking
over several cycles due to sintering of the catalyst. For Ni/TiO2, and steam regeneration, as shown in Fig. 16. Aiello et al. [57]
Ni/Al2O3, and PdeNi(1:3)/SiO2, the catalysts showed good found only trace amounts of nickel oxide on the regenerated
stability over the cycles. samples. SEM micrographs indicated that carbon was

Fig. 14 e a) Fresh catalyst before calcination, and b) after complete regeneration in air [67].
2918 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

the catalyst after regeneration was 0.044 g of carbon per gram


of catalyst, so around 95% of the deposited carbon was
removed by steam gasification.
Muradov et al. [66] used activated carbon as a catalyst
for methane cracking in a fluidized bed. They used steam
to regenerate the deactivated catalyst at 950  C for 0.5 h. By
conducting methane cracking/steam gasification cycles,
Muradov et al. [66] found that the catalyst could regain its
activity completely after regeneration with steam.

6.3. CO2 regeneration

Although steam regeneration and air regeneration are widely


mentioned in the literature as regenerating methods, CO2
regeneration has also been evaluated [71], with the reaction
Fig. 15 e Methane conversion obtained during successive shown in the following equation [5]:
cracking cycles (reprinted from [57], with permission from
Elsevier). C þ CO2 42CO; DH0800  C ¼ 174:5 kJ=mol (16)

Takenaka et al. [72] studied methane cracking over nickel


supported on SiO2, TiO2, and Al2O3 at 550  C and subsequent
removed during steam regeneration, but again small pockets regeneration of the catalyst using CO2 at 650  C. They
of carbon remained, which resisted steam regeneration. observed the production of CO during regeneration, and found
Choudhary et al. [70] studied methane cracking using that 95% of the carbon deposited on the catalyst was con-
nickel supported on different metal oxides and zeolites. verted to CO. After conducting a number of cracking/regen-
Cracking and regeneration were conducted at 500  C using eration cycles for each catalyst, the results showed that for
two parallel fixed beds operated in a cyclic mode and alter- Ni/SiO2 the catalytic activity decreased significantly after the
nating the feed from methane to steam. The total amount of fifth cycle to reach almost half of the initial activity.
carbon deposited on the catalyst was 0.82 g of carbon per gram For example, 10% Ni/SiO2 showed an initial activity of 2000
of catalyst after cracking while the total carbon remaining on (H2/Ni). In the fifth cycle the ratio of H2/Ni decreased to 1000.
The authors attributed the activity loss to sintering. In the
case of Ni/TiO2 and Ni/Al2O3, the catalysts kept their stability
over the six cycles that were conducted.
Pinilla et al. [71] used CO2 to regenerate a deactivated
carbon catalyst but the high temperature required to effec-
tively eliminate all of the residual carbon affected the textural
parameters of the catalyst. After the third regeneration/reac-
tion cycle, they concluded that most of the mass of the cata-
lyst originates from carbon formed during cracking, as the
activated carbon that comprised the original catalyst
completely gasified. Abbas and Wan [73] also used CO2 to
regenerate activated carbon. They conducted methane
cracking at 850 and 950  C. The regeneration experiments
were conducted at 900, 950 and 1000  C. The results showed
that the combination of 950  C as a cracking temperature and
1000  C as the regeneration temperature maintained the
catalytic activity of the activated carbon, with no remarkable
loss of activity for 6 cracking/regeneration cycles.

7. Reaction rate equations and reaction


mechanisms

In this section, different kinetic models that have been


proposed in the literature for methane catalytic cracking are
reviewed. The models range from those based on a detailed
reaction mechanism to a global rate expression, from those
Fig. 16 e XRD patterns of the Ni/SiO2 catalyst after (a) 1, that cover just the stable catalytic activity region to models
(b) 2, (c) 3, (d) 5, and (e) 10 successive cracking/regeneration that describe the whole reaction process including deactiva-
cycles (reprinted from [57], with permission from Elsevier). tion, and from models that calculate methane cracking rates
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2919

to models that give the total amount of carbon deposited.


Table 2 e Mechanism of methane cracking starting with
Overall, the rate equations for methane cracking available in dissociative methane adsorption.
the literature can be divided into three main categories;
CH4 þ I (Vacant site) ¼ CH3 (ad) þ H (ad) (1)
detailed mechanistic rates, interpolated mathematical equa-
CH3 (ad) ¼ CH2 (ad) þ H (ad) (2)
tions, and finally a global rate equation. CH2 (ad) ¼ CH (ad) þ H (ad) (3)
CH (ad) ¼ C (ad) þ H (ad) (4)
7.1. Detailed mechanism rate equations C (ad) ¼ C (dissolved) (5)
2H (ad) ¼ H2 þ 2I (6)

In 1965, Grabke [74] studied methane cracking over g-iron


in the 800e1040  C temperature range. Grabke developed
the first rate equation for methane catalytic dissociation in the So, two additional reaction steps were needed to represent the
form of a reversible reaction rate. Grabke proposed that the diffusion of adsorbed carbon through and around nickel to
1=2
forward reaction is proportional to PCH4 PH2 while the reverse form carbon filaments [25,79] and a third step was needed to
3=2
reaction is proportional to PH2 and the concentration of account for encapsulating carbon formation [80,81].
carbon deposited on the catalyst at low concentrations. In his In the following section each model will be discussed. A
conclusions, Grabke [74] proposed that the methane cracking comparative summary between the different detailed mech-
reaction is a multi-step reaction. Grabke suggested that anism rate models is illustrated in Table 4.
methane undergoes a series of dehydrogenation reactions and
concluded that the rate-controlling step of methane cracking 7.1.1. Models assuming non-dissociative adsorption of
is the formation of the methyl group from methane. methane (Table 1)
Five years later, Grabke [75] continued his research on
methane cracking; also using iron again, but here supported on 7.1.1.1. Grabke [75] model. Assuming that step 3 in Table 1 is
g-alumina as the catalyst and developed a detailed mecha- the rate-limiting step, Grabke [75] studied the kinetics of
nism of the reaction. Grabke [75] suggested a reaction mech- carbon deposition on a g-Fe surface exposed to CH4 and H2 gas
anism that later became the basis for all methane cracking mixtures. Grabke [75] used two iron foils placed inside
mechanisms found in the literature. Grabke concluded that aluminum tubes in the same furnace; one foil was subjected
after methane adsorption on the catalyst, the methyl group to a mixture of CH4 and H2 for carburization (increasing the
passes through a series of dehydrogenation steps, until it ends carbon content of the iron foil using a mixture of CH4 and H2)
up as dissolved carbon. Despite the fact that all the reaction and hydrogen for decarburization (decreasing the carbon
mechanisms proposed in the literature since stem from the content of the iron foil using hydrogen). The other iron foil
same designed by Grabke [74,75], there have been slight was kept in a hydrogen atmosphere as a reference. The elec-
differences proposed, either in the methane dissociation steps trical resistance of the reacting foil was recorded to study the
or by adding more steps to the Grabke mechanism. change in the composition relevant to the reference sample.
Grabke [75] assumed that the adsorption of methane on the The reaction rate of the limiting step was taken as the rate of
catalyst is non-dissociative, as shown in Table 1, which has carburization while the rate of decarburization was used as
been used as the first step for methane cracking mechanism the reverse reaction rate. As a result, the reaction rate is equal
by many researchers [25,76e78]. Although the non-dissocia- to the carburization rate minus the decarburization rate. As
tive adsorption of methane is more often used, dissociative mentioned above, the rate of the forward reaction was found
1=2
methane adsorption has also been proposed [79,80], as shown proportional to PCH4 =PH2 and the reverse reaction rate was
3=2
in Table 2. proportional to PH2 . The methane cracking rate equation
In Grabke’s suggested mechanism, Grabke [75] didn’t (Equation (17)) can be expressed as a dependence on the
explain how the deposited carbon will form carbon filaments. partial pressures of methane and hydrogen as follows:
Grabke proposed carbon adsorption on the active site and
dCCH4
dissolution in the active site, which may explain the deacti-  1=2
¼ kPCH4 =PH  k0 PH
3=2
aC (17)
dt 2 2
vation behaviour, but the mechanism doesn’t reflect the
complete process of forming carbon filaments. The adsorbed where k and k0 are the forward and reverse specific reaction
carbon is an intermediate compound that later forms either constants for the rate-limiting step and aC is the carbon
carbon filaments or encapsulating carbon as shown in Table 3. chemical activity.

Table 1 e Mechanism of methane cracking starting with


non-dissociative methane adsorption.
Table 3 e Carbon diffusion and encapsulating carbon
CH4 þ I (Vacant site) ¼ CH4 (ad) (1)
formation.
CH4 (ad) ¼ CH3 (ad) þ H (ad) (2)
CH3 (ad) ¼ CH2 (ad) þ H (ad) (3) Diffusion in Ni phase: C (dissolved) ¼ C (nickel rear) (1)
CH2 (ad) ¼ CH (ad) þ H (ad) (4) Precipitation from Ni phase: C (nickel rear) ¼ C (2)
CH (ad) ¼ C (ad) þ H (ad) (5) (Carbon filaments)
C (ad) ¼ C (dissolved) (6) Formation of encapsulating carbon: C (ad) ¼ C (3)
2H (ad) ¼ H2 þ 2I (7) (encapsulating)
2920 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

Table 4 e Comparison of different detailed mechanism rate models.


Model Rate-limiting step Catalyst used Temperature Reactor type
range,  C

Models assuming non-dissociative adsorption of methane (Table 1)


Grabke [75] CH3 (ad) ¼ CH2 (ad) þ H (ad) gFe 1000e1150 Tube
Bernardo et al. [82] CH4 þ I (Vacant site) ¼ CH4 (ad) CueNi/SiO2 525e675 Microbalance
Demicheli et al. [76] CH4 þ I (Vacant site) ¼ CH4 (ad) Ni/Al2O3eCaO 565e665 Microbalance
Snoeck et al. [25] CH4 (ad) ¼ CH3 (ad) þ H (ad) Commercial nickel catalyst 500e550 Microbalance
Kuvshinov et al. [77] CH4 þ I (Vacant site) ¼ CH4 (ad) Highly loaded nickel catalyst 530e590 Fluidized bed

Models assuming dissociative adsorption of methane (Table 2)


Alstrup and Tavares [79] 1. CH4 þ I (Vacant site) ¼ CH3 (ad) þ H (ad) Ni/SiO2 450e550 Microbalance
2. CH3 (ad) ¼ CH2 (ad) þ H (ad)
Zavarukhin and CH4 þ I (Vacant site) ¼ CH3 (ad) þ H (ad) 90 wt% NieAl2O3 490e590 Fluidized bed
Kuvshinov [78]
Hazra et al. [80] All steps in Tables 2 and 3 are taken 5% Ni/g-Al2O3 500 Microbalance
into consideration
Borghei et al. [83] CH4 þ I (Vacant site) ¼ CH3 (ad) þ H (ad) NieCu/MgO 550e650 Fixed bed

The results showed that the rate of the forward reaction actual rate of carbon deposition at any time (r) is equal to the
was independent of the carbon atoms dissolved on the iron product of (r*) and (a). (r*) can be defined as the maximum
foil surface. The backward reaction was found to be linearly carbon deposition rate at time (t) while (a) is the activity
dependent on the carbon concentration for low carbon coefficient which accounts for the reduction in the maximum
concentrations. The experimental data agreed well with the rate of carbon deposition due to active sites being blocked. The
results predicted from Equation (17). model was able to accurately predict the rate and the deacti-
vation regime for the catalyst under study.
7.1.1.2. Bernardo et al. [82] model. Bernardo et al. [82] used Using the reaction mechanism shown in Table 1, and
a 1% copper/10% nickel alloy supported over silica catalyst to assuming methane adsorption as the rate-limiting step (step 1
study methane catalytic cracking and steam reforming. A in Table 1), Demicheli et al. [76] developed the following
mixture of methane and hydrogen was used and the rate of equations to calculate the maximum rate for carbon deposi-
carbon deposition over the catalyst was calculated. The tion and the catalyst activity:
proposed rate equation, assuming step 1 in Table 1 as the rate-
  n
limiting step, can be written as follows: PH
r ¼ k PCH4  2 1 þ KH P0:5
H2 (19)
KP
dCCH4  .
 ¼ kPCH4  k0 P2H2 PaH2 ½C (18)
dt For activity:
0
where k and k are again the forward and reverse specific  
a ¼ exp  kd PCH4 s=PH2 (20)
reaction rate constants for the rate-limiting step.
The equation as written above is a modified form of the rate
s ¼ t  t (21)
equation developed by Grabke [75]. Two essential differences
are observed between the two. In Bernardo et al.’s [82] equation, so that the total rate at time t:
the rate is independent of carbon activity and an inhibiting
r ¼ r :a (22)
effect by hydrogen is taken into account, in the denominator of
Equation (18), with experimental data demonstrating inhibi- where k ¼ 2:83  108 expð97; 000=ðRTÞÞ and is the specific rate
tion by hydrogen as shown in Fig. 17. Bernardo et al. [82] found constant for the rate of carbon deposition (g/gcat h kPa) [77]; KP:
that for the CueNi alloy, the value of alpha that best equilibrium constant for methane cracking (kPa) [77];
fits Equation (18) to the experimental data was 0.3. KH ¼ 9:883  109 expð108; 300=ðRTÞÞ and is the equilibrium
The rate equation predicted the experimental data well at constant for hydrogen adsorption (kPa0.5) [77]; kd is the
lower hydrogen pressures, but at higher hydrogen pressures, specific rate constant for the deactivation rate (h1); t* is the
a discrepancy between the predicted and experimental rate time at which carbon deposition rate reaches a maximum (h);
was observed due to high carbon gasification rates. and n is the number of active sites participating in the rate-
limiting step, best fit at n ¼ 7 [77].
7.1.1.3. Demicheli et al. [76] model. Demicheli et al. [76]
studied carbon formation from CH4eH2eN2 mixtures using 7.1.1.4. Snoeck et al. [25] model. Snoeck et al. [25] developed
Ni/Al2O3eCaO in the 565e665  C temperature range. The a model for the formation of filamentous carbon via methane
authors implemented a separable kinetics technique for cracking over a commercial nickel catalyst. The experimental
developing a rate equation to predict carbon deposition. The work supporting this model was done in an electro-balance in
authors used two dependent variables to calculate the reac- the 500e550  C temperature range and a pressure between 1.5
tion rate: the first is the rate of carbon deposition without and 10 bar. This model assumes that methane is first adsor-
deactivation (r*) and the second is an activity factor (a). The bed, followed by a series of dehydrogenation reactions leaving
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2921

Fig. 17 e The rate of carbon deposition on CueNi/SiO2, with a) 1% Cu and b) 10% Cu (reprinted from [82], with permission
from Elsevier).

carbon adsorbed on nickel. This carbon will eventually form 7.1.1.5. Kuvshinov et al. [77] model. Kuvshinov et al. [77]
the carbon filament, as shown in Tables 1 and 3. The release of developed a model predicting carbon deposition as a result
the first hydrogen atom from the adsorbed methane is of methane catalytic cracking. To develop their model, they
assumed to be the rate-determining step, step 2 in Table 1. The studied the effect of varying CH4eH2 and temperature on
reaction steps could be arranged based on the physical a highly loaded nickel catalyst in a flow vibro-fluidized bed
process taking place: surface reactions, dissolution/segrega- catalyst micro-reactor. This model calculates the maximum
tion at the gas/nickel side, diffusion of carbon through nickel, carbon deposition while taking into account catalyst deacti-
and precipitation/dissolution of carbon as filaments at the vation. Kuvshinov et al. [77] used the maximum carbon
nickel/support side. The possible surface reaction mecha- deposition rate developed by Demicheli et al. [76], and then
nisms they described for methane cracking on a nickel surface
are shown in Fig. 18.
By assuming the surface concentrations of adsorbed H, CH,
and CH2 are negligible, the following rate equation was
derived assuming all reaction steps are reversible at high
carbon concentration and a small concentration gradient
around the active site:
!
þ 1 2
km KCH4 $ PCH4   $PH2
KM
r¼ !2 (23)
1 3=2
1 þ 00 $PH2 þ KCH4 $PCH4
Kr

where kþ m ¼ 25; 040expð58; 893=ðRTÞÞ and is the specific rate


constant for the rate-limiting step, mol/gcat.h; kCH4 ¼ 0:21
expð567=ðRTÞÞ and is the equilibrium constant for methane
adsorption, J/mol; K*M is the experimental threshold constant;
and K00r ¼ 5:18  107 expð133; 210=ðRTÞÞ.
The model results showed that the surface concentrations
of adsorbed methane and hydrogen are correlated to the
partial pressures of methane and hydrogen. While for the
carbon surface concentration, the results showed that it is not
dependent on hydrogen and/or methane partial pressures.
The model gave excellent prediction of the carbon formation Fig. 18 e Possible methane cracking reaction pathways
rate at 500  C, as shown in Fig. 19. (reprinted from [25], with permission from Elsevier).
2922 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

gas. Using molecular beam studies of methane interacting


with Ni(111), Alstrup and Tavares modified the Grabke model
to start with direct methane dissociative chemisorption as
shown in Table 2.
Using a Langmuir-type model (the occupation of a site is
independent of other sites being occupied), Alstrup and
Tavares developed the model assuming that: 1) all the reac-
tion mechanism steps are in equilibrium except steps 1e4
in Table 2, 2) the surface species are competing for the surface
sites, and 3) a constant carbon coverage on the catalyst
surface.
The authors derived four rate equations; each equation
representing the rate of carbon deposition assuming that one
of the steps described by 1e4 in Table 2 is the rate-limiting
step. By comparing the experimental results with the pre-
dicted results for each case, Alstrup and Tavares concluded
that assuming steps 3 or 4 as the rate-limiting step didn’t truly
predict the experimental data. While assuming steps 1 or 2 as
Fig. 19 e Model prediction and experimental results of the the limiting steps, the predicted values agreed well with the
carbon formation rate at 500  C using a commercial nickel experimental data, as shown in Fig. 20. Then they modified
catalyst (reprinted from [25], with permission from their model assumptions: steps 1 and 2 are not in equilibrium
Elsevier). and the rates for both steps are equal and the two rates also
equal the overall rate.
Considering step (1) as the rate-limiting step, they devel-
oped the following rate equation:
developed an equation that predicts the maximum amount of
 
carbon deposition until the catalyst is completely deactivated. ICH IH
r1 ¼ k1 PCH4 I2  3 (26)
The rate of carbon formation (g/gcat.h) can be calculated K1
from the following equation:
Considering step (2) as the rate-limiting step, they devel-

n=ðnþ1Þ oped the following rate equation:
n þ 1  1=nþ2  2 
r¼  kr C  C2max þ r1þ1=n (24)  
2 ICH IH
r2 ¼ k2 ICH3 Iv  2 (27)
The maximum amount of carbon deposited (g/gcat) until the K2
catalyst is deactivated is:

1=2
2
Cmax b ¼  þ C2max (25)
k ðn þ 1Þr1=nþ2

where r* is the rate of carbon formation (g/gcat.h), as calculated


from Equation (19) (Demicheli et al. [76] model); Cmax is the
specific weight of carbon formed during time t where the rate
of formation is maximum (g/gcat); n is the number of active
sites participated in the limiting step, best fit at n ¼ 7; C is the
specific weight of carbon formed on the catalyst (g/gcat); and
k ¼ 2:73  1013 expð99; 270=ðRTÞÞ, which is the deactivation
rate constant in (gcat)3. h/g3.
In their conclusions, Kuvshinov et al. [77] found that the
methane catalytic cracking mechanism can be interpreted
using the LangmuireHinshelwood approach. The controlling
step is the dissociative adsorption of methane. Catalyst
deactivation was proportional to the amount of carbon
deposited on the catalyst.

7.1.2. Models assuming dissociative adsorption of methane


(Table 2)

7.1.2.1. Alstrup and Tavares [79] model. Starting with the


Grabke [75] mechanism, Alstrup and Tavares [79] developed
a model for methane catalytic cracking to calculate the carbon Fig. 20 e Predicted and experimental rate of carbon
deposition rate. In their experimental study, they used deposition (reprinted from [79], with permission from
a Ni/SiO2 catalyst and a mixture of CH4 and H2 as the reactant Elsevier).
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2923

where K1 and K2 are the equilibrium constants for reaction


steps (1) to (2) and k1 and k2 are the forward rate constants for
reaction steps (1) and (2), respectively. I is the concentration of
vacant sites. The site balance for the vacant sites could be
expressed as:

I ¼ 1  IH  ICH  ICH2  ICH3  IC (28)

where Ix is the concentrations of sites occupied by species x.

7.1.2.2. Zavarukhin and Kuvshinov [78] model. Zavarukhin


and Kuvshinov [78] developed a mathematical model for the
formation of carbon from methane catalytic cracking using
different mixtures of CH4eH2 and a 90 wt% Ni-Al2O3 catalyst.
Zavarukhin and Kuvshinov [78] used the reaction mechanism
shown in Table 2. The hydrogen volume in the mixture was
varied between 0 and 40%, and the temperature was varied Fig. 21 e Experimental (points) and calculated (line) carbon
between 490 and 590  C in a fluidized catalyst bed micro- deposition amounts as a function of time at 550  C
reactor. The rate-limiting step was assumed to be the release (reprinted from [78], with permission from Elsevier).
of the first hydrogen atom, as shown in step 1 of Table 2.
The maximum rate for carbon formation, g/h.gcat, can be
calculated from the following equation: the differential equations without assuming any rate-limiting
!, step, the model can predict the reaction rate and the change in
P2H  2
rmax ¼ k PCH4  2 1 þ KH P0:5
H2 (29) catalyst weight from the onset of the reaction to complete
Kp
deactivation of the catalyst as a function of time, as shown in
Fig. 22. The deactivation predicted in the model occurs via
The change in total carbon deposited and rate of deposition
a decrease in the number of active sites available for the
can be calculated by integrating the following equations:
reaction due to encapsulating carbon deposition. Hazra et al.
dc [80] also reported that increasing the total pressure may
¼ rmax a (30)
dt increase the catalytic activity of the catalyst over time. They
attributed this pressure increase effect to the increase in
da
¼ ka r2max ca (31) hydrogen partial pressure, which may decrease the rate of
dt
carbon formation and catalyst deactivation at the same time.
where C is the specific weight of carbon deposited on the
catalyst (g/gcat); k ¼ expð20:492  ð104; 200=RTÞÞ and is the 7.1.2.4. Borghei et al. [83] model. Borghei et al. [83] developed
specific rate constant for the rate of carbon deposition (kPa); a kinetic model for methane cracking over Ni-Cu/MgO in the
KP ¼ 5:088  105 expð91; 200=RTÞ and is the equilibrium temperature range 550e650  C using a CH4eH2 mixture at
constant for methane cracking (kPa); KH ¼ expð163; 200= atmospheric pressure. The rate-limiting step is the dissocia-
RT  22:426Þ and is the equilibrium constant for hydrogen tive adsorption of methane, step 1 in Table 2. They developed
adsorption (kPa0.5); kd is the specific rate constant for the the following equation for calculating the maximum methane
deactivation rate (h1); ka ¼ expðð135; 600=RTÞ  32:077Þ and is decomposition rate, in mol/gcat.h, assuming the surface
the deactivation rate constant (gcat)3. h/g3; and a is the activity concentrations of the reaction intermediates are negligible:
coefficient, a ¼ r/rmax.
k 2
Zavarukhin and Kuvshinov [78] applied their model equa- kþ PCH4  P
kr H2
tions to predict the specific carbon content on the catalyst at r ¼ !2 (32)
k 3=2
823  C. The discrepancy between the calculated and experi- 1 þ PH2
kr
mental data didn’t exceed 10%, as shown in Fig. 21.

7.1.2.3. Hazra et al. [80] model. Hazra et al. [80] performed an


experimental study using a thermo-balance at 500  C at 1e10
bar and a 5% Ni/g-Al2O3 catalyst. The authors assumed that
the cracking rate is controlled by the deposition of carbon and
the formation of carbon whiskers. They assumed that deac-
tivation takes place due to the formation of encapsulating
carbon. The 9-step mechanism shown in Tables 2 and 3 was
used. Using the LangmuireHinshelwood mechanism, they
developed a reaction rate equation based on non-separable
kinetics. The model is a series of differential equations that Fig. 22 e Experimental and calculated change in specific
measures the concentration of the reaction intermediates carbon weight at 500  C (reprinted from [80], with
starting with the partial pressures of CH4 and H2. By solving permission from John Wiley & Sons).
2924 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

They defined an activity term, a, as a function of time, which


Table 6 e The effect of reduction temperature on the
can be calculated from the following equation: kinetic parameters in Ref. [10] ([10], with permission from
Elsevier).
1
a¼ 1=ðd1Þ
(33) Tred., CCm jC1  102 jC2 rcW
½1 þ ðd  1Þrd t 
C (g.c/g.cat.) (g.c/g.cat.min) (g.c/g.cat.min) (g.c/g.
cat.min)
so that the overall reaction rate is:
600 0.39801 0.02872 0.16627 4.2
r ¼ r :a (34) 700 0.23620 0.06434 0.32012 6.4
800 0.18625 0.11645 0.55713 6.6
where kþ ¼ 597:98  ð51; 234:6=RTÞ and is the rate constant of
the forward reaction of the rate-limiting step, mol/gcat/h/atm;
k ¼ 17  ð9217:5=RTÞ and is the rate constant of the reverse
reaction of the rate-limiting step, mol/gcat/h/atm1/2; kr ¼ 7.2.1. Catón et al. [10] model
0:607  ð8045:6=RTÞ, atm3/2; d is the order of deactivation, and This model was developed based on an experimental study
the optimum value is 2.11; kd is the specific constant of that was conducted using a Ni/Al2O3 catalyst and a thermo-
deactivation; and rd is a function of temperature, and balance, in the 600e800  C temperature range. Several
methane and hydrogen partial pressures. parameters were studied experimentally to determine their
The deactivation of the catalyst was found to be a second effect on the reaction rate, including hydrogen partial pres-
order reaction. The deactivation is a function of time, sure, methane partial pressure, reaction temperature, and
temperature, and partial pressures of methane and hydrogen. pre-reduction temperature. Deactivation was assumed to be
The model showed excellent agreement with the experi- due to encapsulation of Ni with coke and/or space limitations.
mental data. The driving force for coking was assumed to be the difference
in chemical potential between the gas phase and the
carbon filaments. A maximum number of carbon filaments
7.2. Mathematically fit rate equations can be formed during the nucleation step according to the
reaction conditions, CCm, which consequently determines the
This group of models predicts the total carbon deposited on maximum amount of coke that can be deposited, CC. Based on
the catalyst, but doesn’t take into account the different reac- the previous assumptions, Catón et al. [10] developed the
tion mechanisms that have been proposed for methane following equations.
cracking. Catón et al. [10] and Villacampa et al. [13] divided the The rate of carbon formation during the nucleation stage
reaction term from the onset of reaction to catalyst deacti- can be expressed as:
vation into two physical stages: 1) nucleation and 2) filament
growth, which takes into account the catalyst deactivation dCCN
rCN ¼ ¼ jC1 ðCCm  CC Þ2 þjC2 CC ðCCm  CC Þ (36)
stage. An equation was developed for each stage to calculate dt
the carbon deposited. Then by adding the two amounts The rate of filament growth during the final steady-state
together, the total amount of carbon formed can be calculated period is:
from which the total carbon formation rate can be predicted.
The proposed equations have been fit mathematically rcW ¼ dCCW =dt (37)
(regression analysis) using different parameters. The param- By integrating Equations (36) and (37) and replacing the
eters were correlated to different reaction conditions, which terms in Equation (35), the total carbon deposited is:
are listed in Tables 5 and 6.
For both models, four variables must be defined before jC1 ð1  expð  jC2 CCm tÞÞ
CC ¼ CCm þ rCW t (38)
discussing any details: CCm: the total number of carbon fila- ðjC1 þ ðjC2  jC1 Þexpð  jC2 CCm tÞÞ2
ments that can be formed in the nucleation stage at certain And the overall rate of carbon formation will be:
conditions; CCN: total amount of carbon deposited during the
" #
nucleation stage, mg; CCW: total amount of carbon deposited jC1 ðjC2 CCm Þ2 ð1  expð  jC2 CCm tÞÞ
rc ¼ þ rCW (39)
during the carbon filament stage, mg; and CC: the total content ðjC1 þ ðjC2  jC1 Þexpð  jC2 CCm tÞÞ2
of carbon, mg, where
CCm, jC1 , jC2 (which are fitting parameters), and rcW were
CC ¼ CCN þ CCW (35) correlated using nonlinear regression, and their values at

Table 5 e The effect of reaction temperature on the kinetic parameters discussed in Ref. [10] ([10], with permission from
Elsevier).
Treac.,  C CCm (g.c/g.cat.) jC1  102 (g.c/g.cat. min) jC2 (g.c/g.cat.min) rcW (g.c/g.cat.min)

550 0.18256 5.019 0.2592 7.2


600 0.22010 6.214 0.37254 5.7
625 0.21368 7.224 0.47159 3.1
650 0.14463 13.139 1.57469 4
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2925

different reaction and reduction temperatures are listed in (39) [10] for calculating the total carbon deposited to become in
Tables 5 and 6 [10]. the following form:

jC1 CCm t
7.2.2. Villacampa et al. [13] model CC ¼ CCm þ rCw t (44)
1 þ jC1 CCm t
This model was used to predict the total carbon deposited
from the start of the cracking reaction until complete catalyst
7.3. Global rate equation models
deactivation. A comprehensive experimental study was con-
ducted using a 30% Ni/Al2O3 catalyst in a thermo-balance. The
7.3.1. Fukada et al. [30] rate equation
study parameters were reaction and reduction temperatures
The model developed by Fukada et al. [30] was used to study
and feed composition. This model follows the same trends
the overall cracking rate of methane in a fixed-bed reactor
proposed by Catón et al. [10] and it uses the same equations
using a nickel/silica catalyst. The model was developed to
that were discussed in the Catón et al. [10] model. However,
predict the cracking rate until the catalyst completely deac-
Villacampa et al. [13] conducted a more detailed experimental
tivated. The authors found that the rate can be expressed
study to better understand the effect of operating conditions
using a single reaction rate equation that is first order in
on the deposited carbon and the rate of deposition, to help
methane. The overall first order reaction rate constant,
achieve a better correlation for the model parameters (CCm,
kdecomp, was expressed as:
jC1, jC2, and rcW). The comparison between the model
prediction and the experimental results is shown in Fig. 23;  
kdecomp ¼ 3:09  10exp  29:5 ½kJ=mol=Rg T s1 (45)
the figure shows excellent agreement between the experi-
mental and model results. 7.3.2. Muradov et al. [66] rate equation
Villacampa et al. [13] reported data for the effect of CH4 and The model developed by Muradov et al. [66] was used to study
H2 partial pressures, reaction temperature, reduction the methane cracking rate over activated carbon and carbon
temperature, and reaction generation cycles on the kinetic black. They found that the apparent reaction order is 0.5 in
parameters (CCm, jC1, jC2, and rcW). Based on these data, the methane for both activated carbon and carbon black, as
following equations were developed to predict the kinetic shown in the following equation:
parameters in the 550e650  C temperature range:
rCH4 ¼ k  P0:5
CH4 (46)
CCm ¼ 0:239T  8:433T2 þ 9:913E  7T3  3:88E  10T4 (40)
The apparent activation energy measured was 160e201 kJ/mol
jC1 ¼ 25:587T þ 0:089T2  1:033E  4T3  3:989E  8T4 (41) for activated carbon and 205e236 kJ/mol for carbon black, in
the temperature range of 600e900  C.

jC2 ¼ 5:427T þ 0:0188T2  2:186E  5T3 þ 8:435E  9T4 (42)

rCw ¼ 27:9116Tþ0:0962T2 1:104E4T3 þ4:216E8T4 (43) 8. Fluidized bed reactor applications for
methane catalytic cracking
where T is the temperature in K units.
Fluidized bed reactors have a wide range of applications in the
The value of jC2 was assigned to zero for the regenerated
chemical, metallurgical and petroleum industries. A fluidized
catalyst, because the catalyst was completely sintered during
bed reactor is particularly suitable for methane catalytic
the regeneration period. Therefore, they developed Equation
cracking since it enables continuous addition or withdrawal of
solid particles. The catalyst can be circulated between the
cracking reactor and the regenerator based on the deactiva-
tion rate and the energy requirement of the unit, since the
catalyst is the heat source for the cracking reactor. Therefore
the ratio of the total solid catalyst amount to the total gas
amount must be high enough to prevent a temperature drop
through the fluidized bed [2,84]. Fluidized bed reactors offer
many advantages for solid-catalyzed reactions over fixed-bed
reactors. Those advantages are: ease of temperature control,
ease of maintaining a constant temperature in the bed due to
vigorous mixing provided by fluidization, enhanced heat and
mass transfer, and low cost of catalyst handling [15,66,71].
Fluidized bed reactors are considered the most promising
reactor for the large-scale catalytic cracking of natural gas to
produce hydrogen-rich gas [84].

8.1. Process economics


Fig. 23 e Experimental and predicted data for different
reacting mixtures at 600  C (reprinted from [13], with Thermal energy calculations indicate that the energy
permission from Elsevier). requirement per mole of hydrogen produced from methane
2926 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

catalytic cracking (37.4 kJ/mol H2) is considerably less than The carbon-selling price was assumed U.S. $300/ton. Muradov
that required for producing one mole of hydrogen from the [7] developed a relationship between the market value of
steam reforming process (63.3 kJ/mol H2) [2,15]. In addition, hydrogen and natural gas for the three plant sizes, as shown
the hydrogen produced from methane catalytic cracking does in Fig. 24. The authors also compared methane catalytic
not contain any CO or CO2, which eliminates further down- cracking and methane steam reforming based on the large
stream separation processes. While the ratio of hydrogen plant size (60 MMscfd). The study involved two cases for
produced per mole of methane fed is lower in methane cata- methane steam reforming, with and without CO2 sequestra-
lytic cracking due to the absence of steam, overall economic tion. The carbon price was assumed to range between U.S.
studies have shown that a fluidized bed catalytic cracking unit $160e460/ton. The results showed that methane catalytic
has cost advantages over steam reforming units [5,6,85]. The cracking in a fluidized bed is more economic than steam
fluidized bed catalytic cracking unit should contain two reforming with CO2 sequestration and is comparable with
fluidized bed reactors; one for cracking and one for regener- steam reforming without CO2 sequestration. If future envi-
ating the catalyst. However, studies have also shown that the ronmental regulations impose CO2 sequestration, then
operating cost for the two fluidized bed reactors for cracking methane catalytic cracking will be the better option even
and regeneration and the solids-circulation system are higher without adding the carbon-selling credits.
than the operating cost for the conventional steam reforming A common conclusion in the reported literature economic
unit [22]. analyses of methane catalytic cracking [17,18,57,65] is that the
In 1987 U.S. dollars, Steinberg and Cheng [85] compared carbon-selling price is the key aspect to make the process
different methods for hydrogen production at an industrial economically attractive. However, the carbon-selling price is
scale of 108 SCF/D of hydrogen gas at 300e600 psig. They a function of the properties of the deposited carbon [65]. The
concluded that methane thermal cracking is the most properties of carbon produced depend on the operating
economical method for hydrogen production. For an energy conditions and the type of catalyst used [15,65]. The key is
amount of 106 Btu (w3.7 kmol of H2), they found that hydrogen therefore to devise a process that maximizes the quality of the
produced from methane cracking will cost U.S. $5.1. In Table 7, carbon and to find an optimum balance between selling the
the cost of hydrogen production for each method is given in carbon produced and gasifying it for catalyst regeneration.
terms of the net hydrogen production cost for 103 SCF, the Carbon can be used as a commodity product or sequestered
credit gained from by-products, and finally the net cost for (or stored) for future uses [84]. For example, carbon can be
producing 106 Btu after including the cost of different opera- used as a substitute for carbon black, fibres, graphite, elec-
tions required for processing hydrogen after production, like trodes in the aluminum industry, composites, and carbon
separation and enriching. All the prices are in U.S. $ and the fillers, in tires and plastics, or can be used as a catalyst support
processes are arranged in terms of the net cost of 106 Btu. [5,16]. Also, filamentous carbon is used as an additive in
Muradov [7] performed an economic study of methane polymers and electronic components. Carbon may also be
catalytic cracking in a fluidized bed catalytic reactor and burned in air (Equation (14)) and/or in CO2 (Equation (16)) to
a fluidized bed heater for catalyst regeneration. The final supply heat to the cracking reactor or gasified with H2O
gaseous product of the industrial unit under study was (Equation (15)) to produce additional hydrogen.
assumed >99% by using a gas separation unit. The analysis Separation of the product carbon filaments is a promising
was done at three levels of production, based on unit size in way for producing high quality carbon nanotubes. Separation
MMscfd (Million Metric Standard Cubic Feet per Day), namely: of the carbon filament from the support is much less of an
small (6 MMscfd), medium (20 MMscfd) and large (60 MMscfd). issue [59] than the removal of the metal (e.g. by acid

Table 7 e Comparison of hydrogen production costs for different processes ([85], with permission from the International
Journal of Hydrogen Energy).
Process Total hydrogen By-product By-product Net cost for
production cost in credit (U.S.$) 106 Btu (U.S. $)
U.S. $ for 103 SCF

Methane cracking 2.29 Carbon 0.65 5.1


Hydrocarb process 5.82 Carbon 4.04 5.52
Steam reforming 2.06 Steam 0.16 5.9
Coal gasification with electricity 4.51 Steam 1.8 8.4
chemical shift (Westinghouse)
Partial oxidation 3.12 Sulphur 0.03 9.6
Steam iron 4.75 Power 1.14 11.21
High temperature steam electrolysis 5.06 Oxygen 0.84 13.12
Texaco gasification 4.35 Sulphur 0.08 13.26
Coal gasification with high 4.43 None 0 13.76
temperature electrolysis
K-T gasification 5.12 Sulphur 0.02 15.84
Water electrolysis 6.57 Oxygen 0.83 17.83
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2927

increased significantly in the two-stage fluidized bed reactor


compared to the other configuration. The catalyst life was
significantly prolonged in the two-stage temperature reactor
since the mismatch between carbon production rate and
carbon diffusion rate at high temperature was advantageously
used. The carbon was produced at a higher rate in the high
temperature stage and still diffused through nickel in the
lower temperature stage.
Carbon black and activated carbon were used extensively
in different studies for catalyzing the methane cracking
reaction in a fluidized bed. Lee et al. [27] used different types of
activated carbon to catalyze methane cracking. They studied
several parameters including reaction temperature, gas
velocity, and particle size. The catalyst showed good activity
at the beginning but rapidly deactivated due to channel
Fig. 24 e The relationship between hydrogen and natural blocking by carbon deposition.
gas selling prices (reprinted from [7], with permission from Muradov et al. [66] used a circulating fluidized bed reactor
the National Renewable Energy Laboratory for the U.S. for methane cracking experiments, using carbon black and
Department of Energy). activated carbon catalysts. The catalyst flowed continuously
between the cracker and regenerator as shown in Fig. 25.
Under the test conditions, in the cracker, both catalysts
treatment), the latter still being an important challenge [65]. experienced an initial rapid drop in activity followed by slow
One way to improve the marketing of the filamentous carbon deactivation, with the carbon black taking longer to deacti-
is to increase the amount of carbon deposited on the catalyst, vate. In terms of regeneration, the authors compared steam
which will reduce the metal concentration to acceptable and carbon dioxide as regeneration agents. The activated
marketing limits [65]. For example, nickel may be considered carbon regained its activity completely after regeneration
an impurity for some applications but considered an “addi- using steam, but the carbon black did not.
tive” for others, such as using this carbon as a support for Dunker et al. [15] also used a carbon black catalyst for
a nickel catalyst or using it for hydrogen storage applications. the methane cracking reaction in a fluidized bed reactor at
temperatures between 810 and 980  C. The results showed
8.2. Studies involving fluidized beds and methane a rapid decrease in reaction rate in the first 50 min followed by
catalytic cracking a period of 1000 min where the rate remained stable, before
slowly decreasing again. Dunker et al. [15] attributed the loss of
Universal Oil Products developed the Hypro process for activity to the reduced catalytic activity of the carbon produced
cracking natural gas to produce high-purity hydrogen [21]. The compared to the initial carbon used and the filling of the micro
Hypro process was conducted in a fluidized bed reactor con- pores and internal cavities in the catalyst by deposited carbon.
taining 7% Ni/Al2O3. The fluidized bed cracker was connected Jang and Cha [43] used Fe/Al2O3 catalyst for methane
to a fluidized bed regenerator [22]. Deactivated catalyst was cracking in a fluidized bed at 700  C. They studied several
regenerated by burning the deposited carbon in air using the
regenerator and the catalyst circulation from the regenerator
to the cracker was used to supply heat to the reaction [21]. A
maximum production rate was achieved at ambient pressure
and a cracking reactor temperature of 870  C [22]. Increasing
the total pressure had a strong detrimental effect on the
reaction rate. For example, at 870  C, the hydrogen in the
effluent was reduced from 97.6% at 1 atm to 75.8% at 17 atm.
Weizhong et al. [86] modified a two-stage fluidized bed
reactor to study methane catalytic cracking using a Ni/Cu/
Al2O3 catalyst. The temperature in each stage was controlled
separately. The temperature at the upper section was varied
between 500 and 850  C, while the lower section temperature
was fixed at 500  C. The high temperature stage was used to
ensure high methane conversion, while the lower tempera-
ture stage was used to boost the effective carbon diffusion
through the metal to form filamentous carbon. Weizhong
et al. [86] conducted the same set of experiments with two Fig. 25 e Hydrogen production by the thermo catalytic
other reactors and compared them to this two-temperature decomposition of natural gas: 1dfluidized bed reactor,
stage fluidized bed reactor; a packed bed reactor and a single- 2dheater, 3dgas separation unit, 4dgrinder (reprinted
stage fluidized bed reactor with uniform temperature. Results from [16], with permission from the International Journal
indicated that the catalyst life and the conversion of methane of Hydrogen Energy).
2928 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

operating parameters and their effect on the hydrogen yield. a maximum yield observed at 552  C [19]. Increasing the reac-
The hydrogen yield increased when increasing the reaction tion temperature increases the initial reaction rate, but
temperature and decreased when increasing the methane deactivation becomes faster at higher reaction temperatures
flow rate. They also reported an interesting result concerning due to rapid carbon deposition, which blocks the active sites [27].
catalyst activity; the catalyst activity could be maintained via
attrition of the deposited carbon. 8.3.2. Gas velocity
Shah et al. [87] fabricated a 0.5% Mo/4.5% Fe/Al2O3 catalyst The gas velocity is a key parameter in fluidized bed operation.
and used it for methane catalytic cracking in a fluidized bed The fluidization quality or the gasesolid contacting pattern is
between 650 and 700  C. They used an intermittent mode of strongly dependent on gas velocity. Usually the superficial
operation in a fluidized bed at high methane flow rates and velocity in the fluidized bed is expressed in terms of a multiple
a fixed bed at low methane flow rates. They concluded that the of the minimum fluidizing velocity [27]. If the gas flow rate
continuous fixed-bed mode with low methane flow rate was is higher than that corresponding to the minimum fluidizing
more effective than the intermittent fluidized bed. Shah et al. velocity, bubbles can form, which reduces the contact between
[87] reported that the bench scale reactor they used in the the gases and catalyst. A lower gas velocity reduces the
experimental work had no heat or mass limitations which throughput of the reactor and will not induce fluidization, thus
explained the better performance in the fixed bed compared to eliminating the positive effects of fluidized bed reactors [27].
the fluidized bed, with a higher contact time also achieved in Lee et al. [27] studied the behaviour of activated carbon for
the fixed bed. methane cracking in a fluidized bed at 850  C. At lower gas
Ammendola et al. [44] studied methane catalytic cracking in velocities, methane conversion was higher than that at higher
a fluidized bed between 700 and 850  C using a Cu/Al2O3 cata- gas velocities, as shown in Fig. 26.
lyst. They studied the effect of methane inlet concentration, Several equations are available to predict the minimum
reaction temperature, and the contact time by monitoring fluidizing velocity, Umf. Geldart [89] proposed the following
the total amount of carbon deposited on the catalyst and the equations to calculate Umf (in m/s):
deactivation time. The results showed that increasing the For particles larger than 100 mm
methane inlet concentration increased the total amount of
m h i
carbon deposited until the methane concentration reached Umf ¼ ð1135:7 þ 0:0408  ArÞ1=2 33:7 (47)
rg d
20%, with any further increase in inlet methane concentration
negatively affecting the amount of carbon deposited. The For particles smaller than 100 mm
deactivation time decreased with increasing methane concen-  0:934
tration. Increasing the reaction temperature had a negative d1:8
p rs  rg g0:934
Umf ¼ (48)
effect on the amount of carbon deposited and deactivation time. 1111m0:87 r0:066
g
Increasing the contact time increased the amount of carbon
deposited, but did not show any effect on the deactivation time. where Ar is the Archimedes number, and is defined as:
Pinilla et al. [88] studied methane decomposition in a fluid-  
d3p rg rs  rg g
ized bed using a Ni/Cu/Al2O3 catalyst between 650 and 800  C. Ar ¼ (49)
Different factors were studied including catalyst particle size, m2
reaction temperature, and the space velocity and its ratio to
the minimum fluidizing velocity. They found that increasing
reaction temperature, space velocity, and the ratio of space
velocity to minimum fluidizing velocity increased the reaction
rate and deactivation rate at the same time. So, optimum
conditions should be selected based on the balance between
hydrogen production and catalyst deactivation. Between 650
and 700  C, high hydrogen production and catalyst deactiva-
tion rates were observed, and the carbon was deposited in the
form of carbon nanotubes. At temperatures higher than 700  C,
the carbon was deposited as encapsulating carbon. High
hydrogen production and lower agglomeration were observed
when the catalyst particle diameters were about 150 mm.

8.3. Factors affecting fluidization of methane cracking

To maximize hydrogen production and process economics,


the effects of operating conditions must be well studied and
quantified. In the following sections, the effects of operating
parameters on methane cracking are discussed.

8.3.1. Reaction temperature Fig. 26 e Methane conversion as a function of gas velocity


The stable operating temperature range for methane cracking using activated carbon at 850  C (reprinted from [27], with
using nickel is narrow, between 500 and w550  C, with permission from Elsevier).
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2929

dp is the particle diameter (m); rs is the particle density Umf under real conditions, two common approaches have
(kg/m3); rg is the density of methane (kg/m3); g is the gravity been proposed in the literature for Umf at high temperature.
acceleration factor (m/s2); and m is the gas viscosity (kg/m s).
There are empirical equations that have been used for 8.3.2.1. Empirical equations for predicting Umf at high
predicting the minimum fluidizing velocity for methane temperature. Plenty of equations are available to predict Umf at
cracking. Muradov [16] used the following equation from [90] high temperature, but a common observation is that each
to calculate the minimum flow rate necessary for fluidizing equation is only applicable for a certain material under certain
the catalyst particles using methane: conditions (e.g. [97]). Generalized equations have been
  proposed for different materials under wider operating
0:005d2P 33m rs  rg rg g conditions but the predictive abilities are not always accurate.
G¼ (50)
j2 ð1  3m Þm Sangeetha et al. [93] developed a generalized equation for Umf
that has an error >50% in some cases. With carbon deposition,
where G is the mass flow rate necessary to initiate fluidization using empirical correlations will be difficult unless it was
(g/cm). developed for a similar experimental setup.
Dunker et al. [15] used the following equation from [91] to
calculate the minimum fluidizing velocity in cm/s: 8.3.2.2. Experimental investigation of Umf at high temperature.
  In this approach, which was proposed by Pinilla et al. [94]
33mf d2P rs  rg g based on the approach developed by Pattipati and Wen [97],
Umf ¼ (51)
180ð1  3m Þm the minimum fluidizing velocity is determined under reaction
conditions (by using catalyst particles that undergo methane
where dP is the particle diameter (cm); 3mf is the void fraction catalytic cracking with carbon deposited on them). The
at minimum fluidization conditions; rs is the particle density authors propose using an inert gas, like nitrogen, and plotting
(g/cm3); rg is the density of methane (g/cm3); g is the gravity the pressure drop across a bed of carbonaceous material,
acceleration factor (cm/s2); j is the catalyst shape factor, 1 for made by using the fresh catalyst in methane cracking, against
spherical particles; and m is the gas viscosity (g/cm s). the fluidizing gas velocity, as shown in Fig. 27. Then, the
The temperature and pressure in the fluidized bed affect velocity calculated experimentally for nitrogen is multiplied
the minimum fluidizing velocity [89]. Yang [92] reported that by a factor to account for the effect of using methane. Pinilla
the effect of temperature and pressure on the minimum et al. [94] determined this factor to be 1.4.
fluidizing velocity is dependent on the particle size. The This approach takes into account the change in the fluid
pressure showed nearly no effect on the minimum fluidizing dynamic properties when the particles increase in diameter
velocity for a fine powder (less than 100 mm, Geldart A parti- and decrease in density due to the deposition of filamentous
cles) but the minimum fluidizing velocity decreases with carbon and due to agglomeration and changing void fraction
increasing pressure for coarser particles [89,92]. As predicted at minimum fluidization conditions [92,94]. The maximum
from Equations (47) to (49), the Umf decreases with increasing theoretical pressure drop was calculated from the following
temperature [89,92], and experimental evidence has shown equation:
that increasing temperature caused a reduction in the
minimum fluidizing velocity for the fine powder. But, for DPmax ¼ W=S (52)
coarse powders, experimental work shows that Umf increases
with increasing temperature [89].
With increasing temperature, different characteristics of
the bed may vary, including bed expansion, gas density, gas
viscosity, and bed void fraction [89,93]. On the other hand,
Geldart [89] also discussed the fluidized particles being sub-
jected to density changes and sintering. It is important to also
consider the effect of carbon deposition when the minimum
fluidizing velocity is calculated [94], since carbon deposition
changes particle density and increases the sintering and
agglomeration between the particles (as observed from
preliminary experiments in our laboratory). So, using Equa-
tions (47)e(51) to predict the minimum fluidizing velocity for
the modified particles (after carbon deposition) may result in
prediction error since the correlations in Equations (47)e(51),
or the correlations available in the literature that have been
derived from Ergun’s equation (e.g. [61,95,96]), are sensitive to
the value of the void fraction at minimum fluidization Fig. 27 e Umf and pressure drop, using nitrogen at 700  C
conditions. But the void fraction at higher temperatures is (reprinted from [94], with permission from the
difficult to predict due to the agglomeration, sintering, and International Journal of Hydrogen Energy) Umf-pb: the
cohesion as a result of carbon deposition and high tempera- minimum fluidizing velocity at the onset of fluidization
ture [89,96,97]. To overcome the complication raised from calculated for the particle. Umf-eb: the minimum fluidizing
experiments at high temperatures and to accurately predict velocity at complete fluidization calculated for the bed.
2930 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

where W is the mass of carbon product, in kg, and S is the


cross sectional area in m2.
For circulating fluidized beds, another velocity value that is
important for scale up is the terminal falling velocity.
The terminal falling velocity of a particle, Ut, is its falling
velocity through a fluid [96,98]. To reduce the catalyst carry
over, the operating velocity should be kept between Umf and
Ut. Muradov et al. [66], evaluated Ut during methane cracking,
and calculated the terminal falling velocity using a method
developed by Gibilaro et al. [98]. Kunii and Levenspiel [96]
mentioned the following equation to calculate Ut in cm/s:
2   31=3
6m rs  rg g7
Ut ¼ Ut 4 5 (53)
r2g

For spherical particles:


2 31
6 18 0:591 7
Ut ¼ 4 2 þ  0:5 5 (54)
dp dp
Fig. 28 e Effect of activated carbon particle size and time-
on-stream on methane cracking at 850  C (reprinted from
For particles with shape factor js between 0.5 and 1:
[27], with permission from Elsevier).
2 31
6 18 2:335  1:744js 7
Ut ¼ 6
4 2 þ  0:5 7
5 (55)

dp dp

2   31=3 Fluidized bed models are widely available in the literature;


rg rs  rg g but before using any model to simulate the performance of the
dp ¼ dp 4 5 (56)
m2 system under study, the model should be reviewed to find the
one that best matches operating conditions. Any fluidized bed
where dp is the particle diameter; rs is the particle density; rg is model has at least one of the following limitations [99]:
the density of methane; m is the gas viscosity; and g is the
gravity acceleration factor.  model predictions are not consistent with the experimental
results;
8.3.3. Particle size  model is valid for a narrow range of conditions;
In fluidization, usually Geldart A (30e100 mm) type particles  at least one of the model parameters is not readily obtain-
are used, since it is well known that Geldart A type particles able, curve fitting for the experimental data is required to
lead to smooth fluidization, give better contact between the get these parameters; and
gas and catalyst, and achieve higher catalyst effectiveness  solving the model equations is difficult.
factors [27,84]. Lee et al. [27] studied the effect of the catalyst
particle size on the quality of methane cracking using acti- In general, to model a fluidized bed reactor properly two
vated carbon at 850  C. The results revealed that the smaller sub-models must be developed to simulate the physical
the particle size, the higher the methane conversion due to and chemical phenomena taking place in the fluidized bed
the higher surface area. But the effect of the particle size reactor simultaneously. The physical phenomena correspond
diminished after time-on-stream, as shown in Fig. 28. The to the changes in the bed hydrodynamics, i.e., properties
authors found that, under their operating conditions, the best of the bubble and emulsion phase, whereas the chemical
fluidization quality and contact efficiency was achieved with phenomena correspond to the chemical changes occurring in
a particle size of 108 mm. each phase [100].

8.4. Fluidized bed modelling 8.4.1. Different patterns of fluidized bed models
The fluidized bed models in the literature are usually classi-
Accurate models are required to predict the process response fied based on the flow type regime inside the fluidized bed.
to a change in any operating parameters. This is also true for The flow regime changes with increasing gas velocity [96]. As
methane catalytic cracking in a fluidized bed, to improve the the gas velocity increases, the flow regime changes between
hydrogen output and hence the process economics. Before the bubbling regime, the slugging regime (occurs only in
discussing the fluidized bed model developed by Muradov reactors with small diameters), the turbulent regime, fast
et al. [66], some general features of fluidized bed models fluidization, and pneumatic transport as shown in Fig. 29. The
available in literature, modelling patterns and the associated two common types of fluidized bed models are the bubbling
physical conditions for each model pattern are discussed. fluidized bed and the turbulent fluidized bed.
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2931

Fig. 29 e Flow patterns in a fluidized bed (reprinted from [101], with permission from John Wiley & Sons).

8.4.1.1. Bubbling fluidized bed. There are also two different 8.4.1.2. Turbulent fluidized bed. Turbulent fluidized bed
categories of models widely used in the literature to simu- models are divided into single-phase or two-phase models.
late the bubbling fluidized bed. They are the two-phase and Single-phase models are represented by a plug flow reactor,
three-phase bubbling fluidized bed models. The three-phase most often the axially dispersed plug flow, while two-phase
models were developed assuming the presence of three models are represented as continuously stirred tank models
phases in the fluidized bed [96], namely, the bubble phase, [66].
the wake-cloud phase, and the dense phase as shown in In some cases, the turbulent fluidized bed and the bubbling
Fig. 30. The two-phase models do not include the wake- fluidized bed models are used to simulate the same case based
cloud phase, assuming that the wake-cloud and bubbling on the physics inside the fluidized bed; for example, Gungor
phases are lumped into one phase. and Eskin [102] used a turbulent two-phase model for the
bottom zone of the reactor and a three-phase bubbling bed
model for the upper zone.

8.4.2. Muradov et al. [66] model


This model was developed to simulate and scale up methane
cracking in a cracking-regeneration fluidized bed system
using a carbon catalyst circulating continuously between the
cracker and the regenerator. Muradov et al. [66] developed two
models for the system under study; one model was developed
to simulate methane cracking in a bubbling fluidized bed and
the other was developed to simulate methane cracking in
a turbulent fluidized bed. The two models captured most of
the physics and chemistry inside the fluidized bed. The rate of
methane cracking is calculated using Equation (46).
The carbon flow rate in this study was calculated assuming
a methane conversion of 38%; the maximum conversion
obtained during experiments using activated carbon in a lab-
scale reactor. The carbon flow rate was calculated based on
the heat requirement of the reactor, not on the deactivation
rate of the catalyst.

Fig. 30 e Schematic of a bubbling fluidized bed of carbon Energy required


Carbon flow rate ¼    (57)
particles (reprinted from [66], with permission from the Cpc Tin;C  Tbed
International Journal of Hydrogen Energy).
2932 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

  needs to be developed. It should withstand the damaging


Energy required ¼ mm xDH þ mm Cpm Tbed  Tin;m (58)
mechanical effect of carbon filaments and the high friction
where Cpc is the carbon specific heat, J/mol/K; mm is the molar imposed in the fluidized bed.
flow rate of methane, mol/s; Cpm is the methane specific heat, Optimum operating conditions need to be realized. Process
J/mol/K; x is the methane conversion; and Tbed,Tin,m,Tin,C are variables like methane/hydrogen ratio, temperature, flow
the fluidized bed, methane inlet, and carbon inlet tempera- rate, and pressure need to be further studied to determine
tures respectively, in Kelvin. optimum operating conditions. Similarly, methane cracking
The residence time was calculated as follows: kinetics, including deactivation, need to be confidently
determined, which will help in the development of an accu-
residence time of methane ¼ reactor volume/methane flow rate process model. Modelling can be used to narrow the
rate and research range, in terms of significant parameters. Further-
residence time of carbon ¼ bed mass/carbon circulation rate. more, with the difficulties associated with experimental work
in fluidized beds, modelling is a less costly method to optimize
Muradov et al. [66] solved the previous equations with the operating conditions and to investigate the best way to
mass conversation equations. The model predictions were improve the process economics.
comparable to the experimental results. For large-scale Also, and a critical aspect for the economics, new methods
production, the dimensions of the cracker and regenerator for carbon filament separation and improving the quality of
fluidized beds that were predicted are comparable with those the carbon deposited during methane cracking need to be
of industrial fluid catalytic cracking reactors used in the developed.
petroleum refining industry.

10. Conclusions

9. Current trends, future research Methane catalytic cracking is a promising process to produce
carbon monoxide-free hydrogen. Different catalysts can be
Currently, methane cracking is not used industrially since it is used for methane cracking, including nickel, copper, iron,
not yet economically competitive with well developed activated carbon, and carbon black. Carbon filaments are
methane steam reforming. But with the currently increasing produced as a result of the reaction product carbon diffusing
demand for CO-free hydrogen, a continuous process for through the metal catalyst particle, and then the diffusing
methane cracking is becoming more attractive. High methane carbon is deposited in a filament form on the nickel/support
conversion and utilization of the produced carbon filaments interface. The driving force for carbon diffusion is the
are essential factors for process economics. A continuous temperature gradient or concentration gradient around the
process of methane catalytic cracking is likely needed to catalyst particle. Catalyst deactivation has been discussed in
compete commercially, but rapid catalyst deactivation is a key detail. Encapsulating carbon and/or reactor space limitations
deterrent. A circulating fluidized bed arrangement that has are the primary deactivation causes. Different regeneration
separate cracking and regeneration units represents a poten- methods have been evaluated. Regeneration with steam or
tial solution. The catalyst would move between the two units air is widely used in the literature, and carbon dioxide
for not only regeneration but also for heat transfer. Also, has also been tested for catalyst regeneration. Regeneration
developing methods for separating carbon filaments from the with steam and partial regeneration using air are effective in
catalyst or investigating new applications for the un-sepa- regaining the initial catalyst activity. The reaction kinetics
rated carbon filaments will further the economic potential. have been discussed, and two different mechanisms are
To help with the economics, an active catalyst that can ach- commonly used in the literature. The two mechanisms start
ieve high conversions, and remain stable is therefore critical. with either dissociative or non-dissociative adsorption of
Based on the challenges that impede methane cracking from methane, then the two mechanisms assume that the methyl
competing with other hydrogen production methods, some group goes through a series of dehydrogenation steps until
potential future research and development directions are carbon is produced as a deposit on the catalyst. The reaction
discussed here. models in the literature have been categorized into models
A durable catalyst for fluidized bed applications needs to be based on a detailed mechanism, models based on mathe-
developed. Different characteristics are required. The catalyst matical fitting of experimental data, and global rate models.
needs to: Unfortunately, there is no consensus on the reactant
reaction orders.
- provide high conversion;
- have excellent thermal and chemical stability;
- have high carbon capacity;
- last a long time (until complete deactivation); Acknowledgements
- withstand attrition; and
- be light weight. The authors wish to thank the Natural Sciences and Engi-
neering Research Council of Canada and the Department of
Along these lines, a suitable catalyst support, which as Chemical Engineering in the National Research Centre of
noted, plays a major role in catalyst stability and strength, Egypt for financial support.
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2933

references hydrogen and filamentous carbon: Part I. Nickel catalysts.


Appl Catal A 2000;201:61e70.
[20] Echegoyen Y, Suelves I, Lázaro MJ, Moliner R, Palacios JM.
Hydrogen production by thermocatalytic decomposition of
[1] Malaika A, Krzyzynska B, Kozlowski M. Catalytic
methane over NieAl and NieCueAl catalysts: effect of
decomposition of methane in the presence of in situ
calcination temperature. J Power Sources 2007;169:150e7.
obtained ethylene as a method of hydrogen production. Int J
[21] Pohlenz J, Scott N. Method for hydrogen production by
Hydrogen Energy 2010;35:7470e5.
catalytic decomposition of a gaseous hydrocarbon stream.
[2] Muradov NZ. How to produce hydrogen from fossil fuels
In: Ed. Universal Oil Prod, U.S. Patent No 3,284,161 (UOP);
without CO2 emission. Int J Hydrogen Energy 1993;18:211e5.
1966.
[3] Abbas HF, Daud WMAW. Hydrogen production by
[22] Kermode RI. Hydrogen: its technology and implications. In:
thermocatalytic decomposition of methane using a fixed
Cox KE, Williamson KD, editors. Production technology,
bed activated carbon in a pilot scale unit: apparent kinetic,
vol. 1. Cleveland: CRC Press; 1977. p. 61e115 [chapter 3].
deactivation and diffusional limitation studies. Int J
[23] Chin SY, Chin Y-H, Amiridis MD. Hydrogen production via
Hydrogen Energy 2010;35:12268e76.
the catalytic cracking of ethane over Ni/SiO2 catalysts.
[4] Luengnaruemitchai A, Osuwan S, Gulari E. Comparative
Appl Catal A 2006;300:8e13.
studies of low-temperature water-gas shift reaction over
[24] Rahman M, Croiset E, Hudgins R. Catalytic decomposition of
Pt/CeO2, Au/CeO2, and Au/Fe2O3 catalysts. Catal Commun
methane for hydrogen production. Top Catal 2006;37:
2003;4:215e21.
137e45.
[5] Otsuka K, Takenaka S, Ohtsuki H. Production of pure
[25] Snoeck JW, Froment GF, Fowles M. Kinetic study of the
hydrogen by cyclic decomposition of methane and
carbon filament formation by methane cracking on a nickel
oxidative elimination of carbon nanofibers on supported-
catalyst. J Catal 1997;169:250e62.
Ni-based catalysts. Appl Catal A 2004;273:113e24.
[26] Figueiredo JL, Órfão JJM, Cunha AF. Hydrogen production via
[6] Poirier MG, Sapundzhiev C. Catalytic decomposition of
methane decomposition on Raney-type catalysts. Int J
natural gas to hydrogen for fuel cell applications. Int J
Hydrogen Energy 2010;35:9795e800.
Hydrogen Energy 1997;22:429e33.
[27] Lee KK, Han GY, Yoon KJ, Lee BK. Thermocatalytic hydrogen
[7] Muradov N. Thermo-catalytic CO2-free production of
production from the methane in a fluidized bed with
hydrogen from hydrocarbon fuels. DOE hydrogen program
activated carbon catalyst. Catal Today 2004;93e95:81e6.
review. Baltimore, Maryland, USA: The National Renewable
[28] Rostrup-Nielsen JR. Equilibria of decomposition reactions of
Energy Laboratory for the U.S. Department of Energy; 2001.
carbon monoxide and methane over nickel catalysts. J Catal
p. 271e96.
1972;27:343e56.
[8] ABS-energy-research. The hydrogen economy hydrogen
[29] Alstrup I. A new model explaining carbon filament growth
and fuel cells Ed1; 2006.
on nickel, iron, and NieCu alloy catalysts. J Catal 1988;109:
[9] Ermakova MA, Ermakov DY. Ni/SiO2 and Fe/SiO2 catalysts
241e51.
for production of hydrogen and filamentous carbon via
[30] Fukada S, Nakamura N, Monden J, Nishikawa M.
methane decomposition. Catal Today 2002;77:225e35.
Experimental study of cracking methane by Ni/SiO2
[10] Catón N, Villacampa JI, Royo C, Romeo E, Monzón A. In:
catalyst. J Nucl Mater 2004;329e333:1365e9.
Spivey JJ, Roberts GW, Davis BH, editors. Hydrogen
[31] Rostrup-Nielsen J, Trimm DL. Mechanisms of carbon
production by catalytic cracking of methane using NieAl2O3
formation on nickel-containing catalysts. J Catal 1977;48:
catalysts. Influence of the operating conditions. Stud Surf
155e65.
Sci Catal; 2001:391e8. Elsevier.
[32] Yang RT, Chen JP. Mechanism of carbon filament growth on
[11] Bartholomew CH, Robert JF. Fundamentals of industrial
metal catalysts. J Catal 1989;115:52e64.
catalytic processes. 2nd ed. New Jersey: Wiley-AIChE; 2005.
[33] Snoeck JW, Froment GF, Fowles M. Filamentous carbon
[12] Rodat S, Abanades S, Sans J-L, Flamant G. A pilot-scale solar
formation and gasification: thermodynamics, driving force,
reactor for the production of hydrogen and carbon black from
nucleation, and steady-state growth. J Catal 1997;169:240e9.
methane splitting. Int J Hydrogen Energy 2010;35:7748e58.
[34] Ginsburg JM, Pina J, El Solh T, de Lasa HI. Coke formation
[13] Villacampa JI, Royo C, Romeo E, Montoya JA, Del Angel P,
over a nickel catalyst under methane dry reforming
Monzón A. Catalytic decomposition of methane over
conditions: thermodynamic and kinetic models. Ind Eng
NieAl2O3 coprecipitated catalysts: reaction and
Chem Res 2005;44:4846e54.
regeneration studies. Appl Catal A 2003;252:363e83.
[35] De Bokx PK, Kock AJHM, Boellaard E, Klop W, Geus JW. The
[14] Takenaka S, Ogihara H, Yamanaka I, Otsuka K.
formation of filamentous carbon on iron and nickel
Characterization of silica-supported Ni catalysts effective
catalysts: I. Thermodynamics. J Catal 1985;96:454e67.
for methane decomposition by Ni K-edge XAFS.
[36] Zhang Y, Smith KJ. Carbon formation thresholds and
J Synchrotron Radiat 2001;8:587e9.
catalyst deactivation during CH4 decomposition on
[15] Dunker AM, Kumar S, Mulawa PA. Production of hydrogen
supported Co and Ni catalysts. Catal Lett 2004;95:7e12.
by thermal decomposition of methane in a fluidized-bed
[37] Zhang T, Amiridis MD. Hydrogen production via the direct
reactor e effects of catalyst, temperature, and residence
cracking of methane over silica-supported nickel catalysts.
time. Int J Hydrogen Energy 2006;31:473e84.
Appl Catal A 1998;167:161e72.
[16] Muradov N. Hydrogen via methane decomposition: an
[38] Avdeeva LB, Reshetenko TV, Ismagilov ZR, Likholobov VA.
application for decarbonization of fossil fuels. Int J
Iron-containing catalysts of methane decomposition:
Hydrogen Energy 2001;26:1165e75.
accumulation of filamentous carbon. Appl Catal A 2002;228:
[17] Abbas HF, Wan Daud WMA. Hydrogen production by
53e63.
methane decomposition: a review. Int J Hydrogen Energy
[39] Zhang Y, Smith KJ. CH4 decomposition on Co catalysts:
2010;35:1160e90.
effect of temperature, dispersion, and the presence of H2 or
[18] Avdeeva LB, Kochubey DI, Shaikhutdinov SK. Cobalt
CO in the feed. Catal Today 2002;77:257e68.
catalysts of methane decomposition: accumulation of the
[40] Gac W, Denis A, Borowiecki T, Kepinski L. Methane
filamentous carbon. Appl Catal A 1999;177:43e51.
decomposition over NieMgOeAl2O3 catalysts. Appl Catal A
[19] Ermakova MA, Ermakov DY, Kuvshinov GG. Effective
2009;357:236e43.
catalysts for direct cracking of methane to produce
2934 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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5

[41] Venugopal A, Naveen Kumar S, Ashok J, Hari Prasad D, [61] Fogler HS. Elements of chemical reaction engineering.
Durga Kumari V, Prasad KBS, et al. Hydrogen production by 3rd ed. New Jersey: Prentice-Hall; 1999.
catalytic decomposition of methane over Ni/SiO2. Int J [62] Abbas HF, Daud WMAW. An experimental investigation
Hydrogen Energy 2007;32:1782e8. into the CO2 gasification of deactivated activated-carbon
[42] Chesnokov VV, Chichkan AS. Production of hydrogen by catalyst used for methane decomposition to produce
methane catalytic decomposition over NieCueFe/Al2O3 hydrogen. Int J Hydrogen Energy 2010;35:141e50.
catalyst. Int J Hydrogen Energy 2009;34:2979e85. [63] Bartholomew CH. Mechanisms of catalyst deactivation.
[43] Jang HT, Cha WS. Hydrogen production by the Appl Catal A 2001;212:17e60.
thermocatalytic decomposition of methane in a fluidized [64] Ishihara T, Miyashita Y, Iseda H, Taketa Y. Decomposition
bed reactor. Korean J Chem Eng 2007;24:374e7. of methane over Ni/SiO2 catalysts with membrane reactor
[44] Ammendola P, Chirone R, Ruoppolo G, Russo G. Production for the production of hydrogen. Chem Lett 1995;24:93e4.
of hydrogen from thermo-catalytic decomposition of [65] Suelves I, Lázaro MJ, Moliner R, Corbella BM, Palacios JM.
methane in a fluidized bed reactor. Chem Eng J 2009;154: Hydrogen production by thermo catalytic decomposition of
287e94. methane on Ni-based catalysts: influence of operating
[45] Suelves I, Lázaro MJ, Moliner R, Pinilla JL, Cubero H. conditions on catalyst deactivation and carbon
Hydrogen production by methane decarbonization: characteristics. Int J Hydrogen Energy 2005;30:1555e67.
carbonaceous catalysts. Int J Hydrogen Energy 2007;32: [66] Muradov N, Chen Z, Smith F. Fossil hydrogen with reduced
3320e6. CO2 emission: modeling thermocatalytic decomposition of
[46] Li G, Hu L, Hill JM. Comparison of reducibility and stability methane in a fluidized bed of carbon particles. Int J
of alumina-supported Ni catalysts prepared by Hydrogen Energy 2005;30:1149e58.
impregnation and co-precipitation. Appl Catal A 2006;301: [67] Rahman MS. Catalytic decomposition of methane for
16e24. hydrogen production. Master thesis, University of Waterloo;
[47] Takenaka S, Ogihara H, Yamanaka I, Otsuka K. 2004.
Decomposition of methane over supported-Ni catalysts: [68] Koc R, Alper E, Croiset E, Elkamel A. Partial regeneration of
effects of the supports on the catalytic lifetime. Appl Catal A Ni-based catalysts for hydrogen production via methane
2001;217:101e10. cracking. Turk J Chem 2008;32:157e68.
[48] Li Y, Zhang B, Tang X, Xu Y, Shen W. Hydrogen production [69] Beebe TP, Goodman DW, Kay BD. Kinetics of the activated
from methane decomposition over Ni/CeO2 catalysts. dissociative adsorption of methane on the low index planes
Catal Commun 2006;7:380e6. of nickel single crystal surfaces. J Chem Phys 1987;87:
[49] Guo J, Lou H, Zheng X. The deposition of coke from methane 2305e15.
on a Ni/MgAl2O4 catalyst. Carbon 2007;45:1314e21. [70] Choudhary VR, Banerjee S, Rajput AM. Continuous
[50] McCarty JG, Hou PY, Sheridan D, Wise H. Reactivity of production of H2 at low temperature from methane
surface carbon on nickel catalysts: temperature- decomposition over Ni-containing catalyst followed by
programmed surface reaction with hydrogen and water. In: gasification by steam of the carbon on the catalyst in two
Albright LF, Baker RTK, editors. Coke formation on metal parallel reactors operated in cyclic manner. J Catal 2001;198:
surfaces, vol. 202. Washington: American Chemical Society; 136e41.
1982. p. 253e82 [chapter 13]. [71] Pinilla JL, Suelves I, Utrilla R, Gálvez ME, Lázaro MJ,
[51] Baker RTK, Barber MA, Harris PS, Feates FS, Waite RJ. Moliner R. Hydrogen production by thermo-catalytic
Nucleation and growth of carbon deposits from the nickel decomposition of methane: regeneration of active carbons
catalyzed decomposition of acetylene. J Catal 1972;26: using CO2. J Power Sources 2007;169:103e9.
51e62. [72] Takenaka S, Tomikubo Y, Kato E, Otsuka K. Sequential
[52] Kock AJHM, de Bokx PK, Boellaard E, Klop W, Geus JW. production of H2 and CO over supported Ni catalysts. Fuel
The formation of filamentous carbon on iron and nickel 2004;83:47e57.
catalysts: II. Mechanism. J Catal 1985;96:468e80. [73] Abbas HF, Daud WMAW. Thermocatalytic decomposition of
[53] Boellaard E, de Bokx PK, Kock AJHM, Geus JW. The methane for hydrogen production using activated carbon
formation of filamentous carbon on iron and nickel catalyst: regeneration and characterization studies. Int J
catalysts: III. Morphology. J Catal 1985;96:481e90. Hydrogen Energy 2009;34:8034e45.
[54] Tracz E, Scholz R, Borowiecki T. High-resolution electron [74] Grabke HJ. The kinetics of decarburization and
microscopy study of the carbon deposit morphology on carburization of gammaeiron in methaneehydrogen
nickel catalysts. Appl Catal A 1990;66:133e47. mixtures. Physik Chem 1965;69:409e14.
[55] Sacco A, Thacker P, Chang TN, Chiang ATS. The initiation [75] Grabke H. Evidence on the surface concentration of carbon
and growth of filamentous carbon from [alpha]-iron in H2, on gamma iron from the kinetics of the carburization in
CH4, H2O, CO2, and CO gas mixtures. J Catal 1984;85:224e36. CH4eH2. Metall Mater Trans B 1970;1:2972e5.
[56] Schouten FC, Kaleveld EW, Bootsma GA. AES-LEED- [76] Demicheli MC, Ponzi EN, Ferretti OA, Yeramian AA. Kinetics
ellipsometry study of the kinetics of the interaction of of carbon formation from CH4eH2 mixtures on
methane with Ni(110). Surf Sci 1977;63:460e74. nickelealumina catalyst. Chem Eng J 1991;46:129e36.
[57] Aiello R, Fiscus JE, zur Loye H-C, Amiridis MD. Hydrogen [77] Kuvshinov GG, Mogilnykh YI, Kuvshinov DG. Kinetics of
production via the direct cracking of methane over Ni/SiO2: carbon formation from CH4eH2 mixtures over a nickel
catalyst deactivation and regeneration. Appl Catal A 2000; containing catalyst. Catal Today 1998;42:357e60.
192:227e34. [78] Zavarukhin SG, Kuvshinov GG. The kinetic model of
[58] Shah N, Panjala D, Huffman GP. Hydrogen production by formation of nanofibrous carbon from CH4eH2 mixture over
catalytic decomposition of methane. Energy Fuels 2001;15: a high-loaded nickel catalyst with consideration for the
1528e34. catalyst deactivation. Appl Catal A 2004;272:219e27.
[59] Toebes ML, Bitter JH, van Dillen AJ, de Jong KP. Impact of the [79] Alstrup I, Tavares TM. The kinetics of carbon formation
structure and reactivity of nickel particles on the catalytic from CH4 þ H2 on a silica-supported nickel catalyst. J Catal
growth of carbon nanofibers. Catal Today 2002;76:33e42. 1992;135:147e55.
[60] Trimm DL. Catalysts for the control of coking during steam [80] Hazra M, Croiset E, Hudgins RR, Silveston PL, Elkamel A.
reforming. Catal Today 1999;49:3e10. Experimental investigation of the catalytic cracking of
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 6 ( 2 0 1 1 ) 2 9 0 4 e2 9 3 5 2935

methane over a supported Ni catalyst. Can J Chem Eng 2009; [91] Richardson JE. Incipient fluidization and particulate
87:99e105. systems. In: Davidson JF, Harrison D, editors. Fluidization.
[81] Chen D, Lødeng R, Anundskås A, Olsvik O, Holmen A. London: Academic Press; 1971. p. 25e64.
Deactivation during carbon dioxide reforming of methane [92] Yang WC. Fluidization, solids handling, and processing e
over Ni catalyst: microkinetic analysis. Chem Eng Sci 2001; industrial applications. New Jersey: William Andrew
56:1371e9. Publishing/Noyes; 1998.
[82] Bernardo CA, Alstrup I, Rostrup-Nielsen JR. Carbon [93] Sangeetha V, Swathy R, Narayanamurthy N,
deposition and methane steam reforming on silica- Lakshmanan CM, Miranda LR. Minimum fluidization
supported NieCu catalysts. J Catal 1985;96:517e34. velocity at high temperatures based on Geldart powder
[83] Borghei M, Karimzadeh R, Rashidi A, Izadi N. Kinetics of classification. Chem Eng Technol 2000;23:713e9.
methane decomposition to COx-free hydrogen and carbon [94] Pinilla JL, Moliner R, Suelves I, Lázaro MJ, Echegoyen Y,
nanofiber over NieCu/MgO catalyst. Int J Hydrogen Energy Palacios JM. Production of hydrogen and carbon nanofibers
2010;35:9479e88. by thermal decomposition of methane using metal
[84] Chen Z, Elnashaie SSEH. Steady-state modeling and catalysts in a fluidized bed reactor. Int J Hydrogen Energy
bifurcation behavior of circulating fluidized bed membrane 2007;32:4821e9.
reformer-regenerator for the production of hydrogen for [95] Cheremisinoff NP, Cheremisionoff PN. Hydrodynamics of
fuel cells from heptane. Chem Eng Sci 2004;59:3965e79. gasesolid fluidization. Houston: Gulf Publishing; 1984.
[85] Steinberg M, Cheng HC. Modern and prospective [96] Kunii D, Levenspiel O. Fluidization engineering. 2nd ed.
technologies for hydrogen production from fossil fuels. Int J Boston: Butterworth-Heinemann; 1991.
Hydrogen Energy 1989;14:797e820. [97] Pattipati RR, Wen CY. Minimum fluidization velocity at high
[86] Weizhong Q, Tang L, Zhanwen W, Fei W, Zhifei L, Guohua L, temperatures. Ind Eng Chem Process Des Develop 1981;20:
et al. Production of hydrogen and carbon nanotubes from 705e7.
methane decomposition in a two-stage fluidized bed [98] Gibilaro LG, Di Felice R, Waldram SP, Foscolo PU. Generalized
reactor. Appl Catal A 2004;260:223e8. friction factor and drag coefficient correlations for
[87] Shah N, Ma S, Wang Y, Huffman GP. Semi-continuous fluideparticle interactions. Chem Eng Sci 1985;40:1817e23.
hydrogen production from catalytic methane [99] El-Halwagi MM, El-Rifai MA. Mathematical modeling of
decomposition using a fluidized-bed reactor. Int J Hydrogen catalytic fluidized-bed reactors e I. The multistage three-
Energy 2007;32:3315e9. phase model. Chem Eng Sci 1988;43:2477e86.
[88] Pinilla JL, Suelves I, Lázaro MJ, Moliner R, Palacios JM. [100] Jafari R, Sotudeh-Gharebagh R, Mostoufi N. Modular
Parametric study of the decomposition of methane using simulation of fluidized bed reactors. Chem Eng Technol
a NiCu/Al2O3 catalyst in a fluidized bed reactor. Int J 2004;27:123e9.
Hydrogen Energy 2010;35:9801e9. [101] Grace JR. Contacting modes and behaviour classification of
[89] Geldart D. In: Geldart D, editor. Gas fluidization technology. gasesolid and other two-phase suspensions. Can J Chem
New York: John Wiley; 1986. p. 11e32 [chapter 2]. Eng 1986;64:353e63.
[90] Othmer D, editor. Fluidization. New York: Reinhold Publ.; [102] Gungor A, Eskin N. Hydrodynamic modeling of a circulating
1956. fluidized bed. Powder Technol 2007;172:1e13.

Вам также может понравиться