Você está na página 1de 9

Fuel 239 (2019) 896–904

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

A comparison of two-phase and three-phase CO2 methanation reaction T


kinetics
Jonathan Lefebvrea, , Siegfried Bajohra, Thomas Kolba,b

a
Karlsruhe Institute of Technology, Engler-Bunte-Institut, Fuel Technology, Engler-Bunte-Ring 1, 76131 Karlsruhe, Germany
b
DVGW Research Center at Engler-Bunte-Institut, Engler-Bunte-Ring 1, 76131 Karlsruhe, Germany

ARTICLE INFO ABSTRACT

Keywords: In a previous publication related to three-phase CO2 methanation (3PM) reaction kinetics (Lefebvre et al., 2018)
CO2 methanation it was postulated that (i) the liquid phase influences the effective reaction rate but not the intrinsic chemical
Reaction kinetics reaction rate and (ii) gas concentration in the liquid phase, not gas partial pressure, is the relevant parameter to
Three-phase methanation describe 3PM reaction kinetics. In this earlier publication, it was also reported that (iii) measurement un-
Power-to-Gas
certainties related to gas concentration in the liquid phase are high and (iv) catalyst reoxidation during the
starting procedure of the three-phase experiments may not have been fully excluded.
The aim of the present publication is to prove the postulates (i) and (ii). To achieve this, the two-phase CO2
methanation (2PM) reaction kinetics is investigated in a plug flow laboratory reactor. Using the data of 213
validated experiments, a power law kinetic rate equation is developed, which describes 2PM reaction kinetics on
a commercially available catalyst for inlet CO2 partial pressures of 1 bar and temperatures between 200 °C and
300 °C.
This two-phase kinetic rate equation is applied to calculate 3PM reaction rates using temperatures and gas
concentrations in the liquid phase from previous 3PM experiments. It is shown that the two-phase kinetic rate
equation can describe 3PM experiments with good agreement, i.e. a liquid phase does not influence the intrinsic
reaction rate but the concentration of reacting species on the catalyst surface and gas concentration, not gas
partial pressure, is the relevant parameter to describe the CO2 methanation reaction kinetics.

1. Introduction applications in highly efficient final energy conversion technologies,


e.g. decentralized/central combined heat and power units, mobility
The German government has set challenging goals for the German (compressed natural gas and liquefied natural gas) or combined cycle
energy transition “Energiewende”: for the year 2050 a 90% reduction in power plants.
CO2 emissions as compared to 1990 is targeted [1]. This ambitious goal PtG is a flexible process allowing for on-demand energy storage. As
requests a high share of fluctuating renewable energy input which has CO2 methanation is a highly exothermic reaction (see Eq. (1)), an ef-
to be integrated into a stable energy supply system. Therefore, new ficient and flexible heat removal from the reactor is mandatory for the
technologies for energy sectors coupling and long-term energy storage design of an adequate PtG methanation unit [6,7]. Rönsch et al. [8]
are mandatory for the future energy system. showed that state-of-the-art adiabatic fixed-bed methanation reactors
The Power-to-Gas (PtG) process connects the power grid to the are not appropriate for dynamic PtG modes of operations, i.e. new
natural gas grid. Electrical energy is transformed into a chemical energy methanation reactor designs are necessary. A slurry bubble column
carrier: in a first step H2 is produced from electrical power via elec- reactor is an attractive methanation reactor concept for PtG, as this type
trolysis. In a second step CO2 from different sources reacts with H2 to of three-phase reactor grants very efficient reactor temperature control
produce synthetic natural gas (SNG) (see Eq. (1)) [2–5]. as well as good dynamic behavior [9,10]. For the design of a slurry
bubble column reactor, knowledge of reactor hydrodynamics, mass
CO2 (g) + 4 H2 (g) CH 4 (g) + 2 H2 O(g) hRSTP = 165. 1 kJ/mol (1)
transfer as well as reaction kinetics is required [11,12]. The reaction
SNG offers a number of benefits over H2, i.e. a very well-developed kinetics of the CO2 methanation has been investigated extensively in
infrastructure for transportation and storage and a wide range of two-phase (gas/catalyst) systems, while three-phase CO methanation


Corresponding author.
E-mail address: jonathan.lefebvre@kit.edu (J. Lefebvre).

https://doi.org/10.1016/j.fuel.2018.11.051
Received 10 September 2018; Received in revised form 28 October 2018; Accepted 8 November 2018
0016-2361/ © 2018 Elsevier Ltd. All rights reserved.
J. Lefebvre et al. Fuel 239 (2019) 896–904

Nomenclature hRSTP Reaction enthalpy at standard conditions (kJ/mol)


v Molecular volume (–)
Symbols Void fraction (–)
mod,CO2 Modified CO2 residence time (see Eq. (6)) (kg·s/mol)
Bo Bodenstein number (see Eq. (3)) (–)
c Concentration (see Eq. (12)) (mol/m3) Abbreviations
d Diameter (m)
D Diffusion coefficient (m2/s) 2PM Two-phase methanation
EA Activation energy of the reaction (kJ/mol) 3PM Three-phase methanation
k Reaction rate constant (see Eq. (10)) (mol/ CSTR Continuous Stirred-Tank Reactor
(kg·s·mol0.5·m−1.5)) DBT Dibenzyltoluene
k0 Pre-exponential factor in Arrhenius equation (mol/ GC Gas Chromatograph
(kg·s·mol0.5·m−1.5)) PFR Plug Flow Reactor
K Parameter to express the reaction rate limitation due to PtG Power-to-Gas
chemical equilibrium closeness (see Eq. (11)) (–) SBCR Slurry Bubble Column Reactor
K eq Equilibrium constant of the reaction (–) SNG Synthetic Natural Gas
K H2O H2O adsorption constant (m3/mol) STP Standard temperature and pressure (273.15 K and
L Length (m) 1.01325 bar)
m Mass (kg) TGA Thermogravimetric Analysis
M Molar mass (kg/mol)
n Molar flow rate (mol/s) Subscripts and superscripts
p Absolute pressure (bar)
Pe Peclet number (see Eq. (17)) (–) ax Axial
r Catalyst mass-specific reaction rate (see Eq. (8)) (mol/ bed Bed material
(kg·s)) cal Calculated
r̄ Integral catalyst mass-specific reaction rate (see Eq. (7)) cat Catalyst
(mol/(kg·s)) exp Experimental
R Ideal gas constant (J/(mol·K)) G Gas phase
T Temperature (°C or K) i Gas component i
u Superficial velocity (m/s) in Reactor inlet
XCO2 CO2 conversion (see Eq. (5)) (–) L Liquid phase
y Gas molar fraction (–) out Reactor outlet
H2 reaction order (–) P Particle
CO2 reaction order (–) t Tube
H2O reaction order (–) * Gas/liquid equilibrium

has been the topic of several publications [30–39]. However, little re- of the corresponding Henry’s law constants is quite high [13].
search has been done on reaction kinetics of the three-phase CO2 me- From this background, the present publication intends to confirm
thanation (3PM). statements (i) and (ii) using a different experimental approach. If the
In a previous work carried out at Engler-Bunte-Institut, Fuel liquid phase only influences the reaction rate in terms of effective gas
Technology (EBI ceb) a 3PM reaction kinetic was derived from ex- concentrations at the catalyst surface, operating a two-phase fixed-bed
periments in a continuous stirred-tank reactor (CSTR) [13]. Based on reactor under three-phase methanation (3PM) operating conditions
experiments using three different liquids, it was speculated that (i) the should lead to identical reaction rates. These operating conditions are
effective reaction rate depends on the liquid specification but not the marked by sub-stoichiometric H2/CO2 ratios. Indeed, due to the strong
intrinsic chemical reaction rate and (ii) gas concentration in the liquid solubility difference between H2 and CO2 in the applied liquid phases, a
phase, not partial pressure in the gas phase, is the relevant parameter to H2/CO2 ratio of 4 in the gas phase – typical for two-phase methanation
describe the CO2 methanation reaction kinetics. Based on these ex- (2PM) – results in a H2/CO2 ratio of about 1.3 in the liquid phase when
periments, a kinetic rate equation for 3PM described in Eq. (2) has been the gas/liquid phase equilibrium is reached [9,13,14]. Furthermore, for
derived: three-phase methanation the maximum allowed reactor temperature is
limited by the temperature stability of the liquid, which sets the max-
79061 cH0.32,L· cCO
0.1
2,L imum reaction temperature to about 350 °C [9,13,14]. Graaf et al. have
r3PM = 3.90699·105·exp · ·K
R·T (1 + 1· c H2 O,L )
0.1
(2) followed the same path to study the liquid phase influence on three-
phase methanol synthesis kinetics: they measured the methanol
In this publication, several catalyst samples from a single catalyst
synthesis kinetics in a two-phase reactor [15] and then in a three-phase
batch were prepared for the experiments: they had been reduced in a
reactor [16–18]. They showed that the activation energy derived from
fixed-bed reactor (see Section 2.1) and then suspended in different li-
three-phase methanol synthesis experiments is much lower than the
quids under an inert atmosphere before proceeding to the methanation
activation energy derived from two-phase methanol synthesis experi-
experiments [13]. As it cannot be fully excluded that catalyst reoxida-
ments [16]. In contrary to our statement in [13], they concluded that a
tion and deactivation took place during the catalyst suspension proce-
liquid phase influences significantly the reaction kinetics of the me-
dure, there is a need to confirm the statements (i) and (ii) in a reacting
thanol synthesis.
system where catalyst reoxidation is excluded. In addition, it was re-
Recent focus has been put on the reaction kinetics of two-phase CO2
ported previously that a significant part of measurement uncertainty is
methanation for PtG application. Experimental investigations were
related to the calculation of H2 concentration in the liquid phase, which
conducted either in a fixed-bed reactor [19] or in a spinning-basket
depends on the knowledge of Henry’s law constants [13]. As the solu-
reactor [20] with Ni catalysts, for temperatures in the range of
bility of H2 in organic oils is usually low, the measurement uncertainty

897
J. Lefebvre et al. Fuel 239 (2019) 896–904

170–340 °C, pressures from 1 to 20 bar and H2/CO2 ratios from 0.25 to 103
8. These publications as well as earlier publications [21–29] showed (H2/CO2)in = 1

CO2 reaction rates r2PM / mmol/(kg·s)


that H2 and CO2 partial pressures have a positive influence on the CO2 (H2/CO2)in = 2
methanation kinetics, while CH4 shows no influence on the methana- 300 °C (H2/CO2)in = 3
tion reaction kinetics [19,20]. The influence of H2O on the CO2 me- 2
280 °C (H2/CO2)in = 4
thanation reaction kinetics was recently investigated by Lim et al. [20]. 10
260 °C (H2/CO2)in = 5
They showed that H2O has a strong negative influence on the CO2 re-
action kinetics. 240 °C
In this publication kinetic experiments were carried out in a two-
220 °C
phase fixed-bed reactor under three-phase CO2 methanation conditions.
101
Hereby, the influence of temperature and reactant as well as product 200 °C
partial pressures on the 2PM reaction kinetics was studied. Based on
experimental data, a kinetic rate equation was derived and used to
reproduce three-phase CO2 reaction rates obtained from earlier ex-
periments [13]. 100
1,7 1,8 1,9 2,0 2,1 2,2
2. Material and methods Inverse temperature 1000/T / 1/K

Fig. 2. Arrhenius plot: influence of temperature and inlet H2/CO2 ratio on the
2.1. Material CO2 reaction rates ( pR = 9.2 bar, pCO2,in = 1 bar, pCH4,in = pH2O,in = 0 bar,
mod,CO2 = 2 kg·s/mol (open symbols) and mod,CO2 = 10 kg·s/mol (closed sym-
The experiments were carried out in the experimental setup shown bols)).
in Fig. 1, which is similar to the one used by Iglesias et al. [40,41].
The gases (CO2, H2, CH4, H2O, Ar, and N2 with purities higher than
99.995%) were supplied via a set of mass flow controllers (Bronkhorst). vertical position of the reaction zone to one of the heating blocks. The
The gas mixture was heated up to the desired inlet temperature before glass reactor itself was placed in a stainless-steel tube to allow experi-
entering the reactor. ments at high pressures. The annular space between the glass and the
The reactor is a glass tube (dt = 8 mm and Lt = 700 mm) used as metal reactor was sealed with an O-ring to prevent gas recirculation and
plug flow reactor (PFR). The glass tube was filled with catalyst particles bypass effects. The steel tube was heated via three heating blocks which
(dP = 50–100 µm) diluted with SiO2 particles (dP = 100–160 µm) in were insulated with glass wool against the environment. This con-
order to mitigate temperature hot spots within the particle bed. The struction is shown in Fig. 12 in Appendix 1.
weight ratio of catalyst to SiO2 was 1–8. Isothermal conditions along the fixed bed were systematically ver-
The catalyst particle size range was chosen to minimize pressure ified using the moveable thermocouple, i.e. the axial temperature
drop and to rule out inter- or intra-particle mass and heat transfer spread ( Tax ) was below 1 K for all experiments. Isothermal conditions
limitations [42–44]. The absence of intra-particle mass transfer lim- were reached by application of high gas velocities and fine tuning of the
itation was validated by the Weisz-Prater criterion [42] for each kinetic three reactor heating blocks. Two catalyst beds were prepared for the
measurement. The temperature profile along the catalyst bed was experiments reported in this publication: one bed with 0.5 g of catalyst
measured with a moveable thermocouple placed in a centrally posi- and a second bed with 0.1 g of catalyst. Indeed, isothermal condition
tioned thermowell. A bed of inert particles was placed in front of the could not be achieved with the first catalyst bed at temperatures higher
catalyst bed to serve as gas mixing and preheating zone. A second bed than 260 °C. With the second catalyst bed, isothermal conditions were
of inert particles was positioned after the catalyst bed to adjust the reached even for the highest investigated reactor temperature of 300 °C.

FIC
PIRC Pressure
Condensate controller
PI
CO2 tank
FIC

H2
FIC
TIC TC

CH4
FIC TC Off
TIC
N2
gas
FIC

Ar TIC

H 2O
TIC GC
PI
FIC

Analysis of
CO2, H2, CO, CH4
Evaporator Fixed-bed reactor Ar, N2 and C1-2

Fig. 1. Flow scheme of the experimental setup.

898
J. Lefebvre et al. Fuel 239 (2019) 896–904

250 adjusted to maintain a constant overall volume flow rate of 620 ml/min
200 °C (STP) at the reactor inlet. A broad range of H2/CO2 ratio from 1 to 5
220 °C was chosen to mimic the sub-stoichiometric conditions of the three-
CO2 reaction rate r2PM / mmol/(kg·s)

200 240 °C
phase methanation with typical H2/CO2 ratios of 1–2 [9,13], and to also
260 °C
cover the H2/CO2 ratios of conventional two-phase methanation. In
280 °C
addition to the experiments shown in Table 1, experiments were per-
150 300 °C
formed to investigate the influence of inlet CO2 partial pressure on the
CO2 reaction rates. For these experiments, pCO2,in was varied from 0.75
to 1.25 bar resulting in H2/CO2 ratio ranging from 0.8 to 6.6. For both
100
investigated catalyst beds, no deactivation was observed during the
experiments (time on stream higher than 700 h for each catalyst bed).
To check the applicability of PFR behavior on the fixed-bed reactor,
50
the Bodenstein number Bo , which compares advective and diffusive
mass transfer, was calculated for each experimental condition ac-
0 cording to Eq. (3). The estimation of the axial Peclet number Peax is
0 1 2 3 4 5 6 described in the Appendix 2.
H2 partial pressure pH ,in / bar Peax ·LR
2 Bo =
dP (3)
Fig. 3. Influence of inlet H2 partial pressure on the CO2 reaction rate for tem-
peratures between 200 and 300 °C ( pR = 9.2 bar, pCO2,in = 1 bar, The Bodenstein number range is 541–1201. Therefore, the PFR as-
pCH4,in = pH2O,in = 0 bar, mod,CO2 = 2 kg·s/mol (open symbols) and sumption (Bo> 100) is verified for all applied experimental conditions
mod,CO2 = 10 kg·s/mol (closed symbols)). [45].

At the reactor outlet, the product gas was cooled to 15 °C to con- 2.2.2. Data interpretation
dense most of the produced water. After passing a pressure controller N2 was used as reference gas to determine the reactor outlet molar
(Bronkhorst), the product gas was analyzed in a gas chromatograph (GC flow rate for each gas component. Knowing the reactor inlet molar flow
7890A, Agilent) for CH4, CO2, CO, Ar, N2, H2, C2H4, and C2H6. rate n in as well as the N2 molar fraction at the reactor inlet and outlet
The catalyst used in this work was a commercially available Ni/SiO2 from gas chromatographic (GC) analysis, the outlet molar flow rate of
catalyst for methanation applications. It was from the same catalyst each gas component can be calculated from the measured gas molar
batch as used previously for the three-phase methanation experiments fraction yi,out according to Eq. (4).
[13]. Prior to each experiment, the catalyst was reduced for 24 h at
yN2,in
400 °C in an Ar/H2 = 1/1 atmosphere flowing at 40 l/h (STP). After the ni,out = nout ·yi,out = n in · ·yi,out
reduction procedure, the reactor was cooled to reaction temperature
yN2,out (4)
and the desired gas atmosphere was applied. The same reduction
During each experiment, attention was paid to the atomic balance
method was applied for the catalyst used in three-phase methanation
over the reactor; measurements with a carbon and hydrogen balance
experiments [13], i.e. the catalyst was reduced in a fixed-bed reactor
error higher than ± 1% were rejected. H2O could not be detected by the
under the above mentioned conditions before being suspended in the
applied GC; the outlet H2O molar flow rate was calculated from the
liquid phase under an inert atmosphere.
oxygen balance over the reactor. The CO2 conversion XCO2 is expressed
Previous three-phase methanation kinetic investigations were con-
with Eq. (5):
ducted with a pre-reduced catalyst and catalyst mass-specific reaction
rates were determined with the reduced and not the oxidized catalyst nCO2,in nCO2,out nCO2,reaction
XCO2 = =
mass. In this publication, oxidized catalyst particles were placed inside nCO2,in nCO2,in (5)
the fixed-bed reactor. Hence, knowledge of the catalyst mass loss after
the reduction procedure was required to compare two-phase and three- With the modified CO2 residence time mod,CO2 , see Eq. (6),
phase CO2 methanation reaction rates. To gain such knowledge, ther- m cat
=
mogravimetric analyses (TGA) were performed in a TGA device, model mod,CO2
nCO2,in (6)
209 F1 provided by Netzsch: three oxidized catalyst samples from 16 to
20 mg were first heated to 400 °C with a ramp of 8 K/min and then the integral CO2 reaction rate r̄2PM can be calculated according to Eq.
reduced for 24 h. For both heating and reduction programs, a volume (7).
flow rate of 50 ml/min (STP) with a composition N2/H2 = 1/1 was sent nCO2,reaction XCO2
through the TGA device. The average catalyst mass loss after reduction r¯2PM = =
mcat (7)
was 22.88%. Using this mass loss, the mass of oxidized catalysts applied
mod,CO2

to the fixed-bed reactor can be converted to the reduced catalyst mass. The true CO2 reaction rate r2PM is calculated solving the PFR design
equation Eq. (8):
2.2. Experimental methods
Table 1
Investigated experimental conditions for the development of a 2PM kinetic rate
2.2.1. Experimental conditions
equation.
For the development of a kinetic rate equation, a total of 213 ex-
periments were conducted under systematic variation of the parameters Parameter Unit Value
listed in Table 1 in order to obtain a broad range of operating condi- T °C 200; 220; 240; 260; 280; 300
tions and CO2 conversion rates (0.02 XCO2 0.75). pH2,in bar 1; 2; 3; 4; 5
All experiments were conducted with an inlet CO2 partial pressure pCH4,in bar 0; 0.4; 0.8
of 1 bar and a reactor absolute pressure of 9.2 bar. N2 was fed at a pH2O,in bar 0; 0.4; 0.8; 1.2; 1.6
constant volume flow rate of 50 ml/min as reference gas for closing the mod,CO2 kg·s/mol 1.6; 2; 2.7; 8; 10; 13
system mass balance. Ar was used as inert buffer gas; the flow rate was

899
J. Lefebvre et al. Fuel 239 (2019) 896–904

XCO2
dXCO2 EA 1
log10k = log10k 0 ·
mod,CO2 = R·ln10 T (13)
0
r2PM (8)

To solve Eq. (8), a kinetic rate equation r2PM must be predicted be-
forehand. 3. Results and discussion
In the following diagrams, the integral reaction rate r̄2PM which is
directly derived from the experiments is used to describe the experi- 3.1. Influence of temperature and gas partial pressure on the CO2 reaction
mental results. The integral CO2 reaction rate is used rather than the rate
CO2 conversion, as it takes into account the CO2 residence time within
the reactor ( mod,CO2 ). This is particularly important when the influence The temperature influence on the integral CO2 reaction rate r̄2PM is
of CO2 partial pressure on the methanation reaction kinetics is in- shown in Fig. 2 for H2/CO2 ratios ranging from 1 to 5 and temperatures
vestigated (see Figs. 4 and 5). Indeed, for these experiments the CO2 ranging from 200 to 300 °C.
partial pressure and the CO2 residence time had to be varied simulta- Between 200 and 300 °C, the CO2 reaction rate is almost doubled for
neously through variation of the inlet volume flow rate of CO2, while each temperature increase of 20 K. For all H2/CO2 ratios, the tem-
keeping the other relevant operating parameters (except N2 volume perature dependence is the same, and apparent activation energies of
flow rate) constant. 73 to 78 kJ/mol can be derived from the experiments. These activation
For each measurement, the data accuracy is evaluated with the energies are typical for the CO2 methanation reaction [19,21,22],
method of partial derivatives. The resulting error bars are given in each which confirms that the experiments were performed in absence of
of the following diagrams. mass and heat transfer limitations [51]. Additionally, Fig. 2 shows that
r̄2PM increases with increasing H2/CO2 ratio. This results from the po-
sitive influence of increasing H2 partial pressure on the CO2 reaction
2.2.3. Development of a reaction rate equation
rates, as shown in Fig. 3.
A power law rate equation with product limitation (see Eq. (9)) as
Fig. 2 also shows that the influence of H2/CO2 ratio is more pro-
reported in [13] was used to describe the CO2 reaction rates observed in
nounced for small H2/CO2 ratios. This effect can be explained by the
this work. Side CO2 reactions to product other than CH4 (e.g. CO) were
higher production of CO at small H2/CO2 ratios: at 240 °C the CO se-
not considered as the selectivity to CH4 was higher than 90% for all the
lectivity is about 1% for H2/CO2 = 1, while it is about 0.4% for higher
experiments.
H2/CO2 ratios. The same trend can be observed for the other in-
cH2 · cCO2 vestigated temperatures. As the presence of even a few ppm of CO is
r2PM = k· ·K known to mitigate the CO2 methanation [21], the higher CO production
(1 + K H2O·c H2O) (9)
for H2/CO2 = 1 leads to stronger mitigation of the CO2 methanation
k is the reaction rate constant as defined in Eq. (10), while K H2O reaction kinetics as compared to higher H2/CO2 ratios.
describes the adsorption constant of H2O on the nickel catalyst. As the The influence of inlet H2 partial pressure pH2,in on the integral CO2
adsorption enthalpy of H2O was not found in literature, K H2O was set to reaction rate r̄2PM is shown in Fig. 3 for temperatures ranging from 200
1 m3/mol in this work, i.e. the influence of H2O on the reaction rate to 300 °C. An increase in pH2,in has a positive influence on r̄2PM, which is
does not vary with temperature (see Section 3.1). Other values were confirmed by several publications [20–24,29]. At 300 °C, r̄2PM is en-
tested for K H2O and delivered similar fits. hanced by 70% when pH2,in is increased from 1 to 4 bar. As previously
reported, the increase in r̄2PM is more pronounced for pH2,in in the range
EA of 1 to 2 bar as compared to higher pH2,in . With a logarithmic linear-
k = k 0·exp
R ·T (10) ization of the experimental data depicted in Fig. 3, the order of reaction
for H2 is determined for each investigated temperature. This order of
K describes the limitation of r2PM when the reaction system ap- reaction increases with increasing temperature from 0.33 to 0.42. In
proaches the chemical equilibrium described by K eq (see Eq. (11)). literature, H2 reaction orders from 0.21 to 1 have been reported
Usually, gas partial pressures are used to derive kinetic rate equation [21–28].
for CO2 methanation [19–26,28,29,46–49]. However, Eq. (9) is written
in terms of gas concentrations, as the principles of 3PM are investigated 250
in this publication. Concentrations ci are estimated with Eq. (12) which 200 °C
is valid under the premise of ideal gas behavior. 220 °C
CO2 reaction rate r2PM / mmol/(kg·s)

200 240 °C
cH22O,out ·cCH4,out c02 260 °C
K=1 ·
cH42,out ·cCO2,out K eq (11) 280 °C
150 300 °C
pi
ci =
R· T (12)
100
The chemical equilibrium constant K eq was estimated through
minimization of the system’s Gibb’s enthalpy. The species’ Gibb’s en-
thalpies were calculated based on the species’ enthalpies and entropies 50
according to the Shomate equation and data from NIST Chemistry
WebBook [50]. The reaction kinetic parameters k , , and were
determined by a least-square minimization of the deviation between 0
0,25 0,75 1,25 1,75
calculated CO2 conversion XCO2,cal (derived from Eq. (8)) and experi-
mentally observed CO2 conversion XCO2,exp (Eq. (5)). As a first step, , CO2 partial pressure pCO ,in / bar
and were guessed and k was determined for each experimental tem-
2

perature. Then, k 0 and EA were calculated from linear regression in an Fig. 4. Influence of inlet CO2 partial pressure on the CO2 reaction rate for
temperatures between 200 and 300 °C ( pR = 9.2 bar, pH2,in = 4 bar,
Arrhenius plot (see Eq. (13)). The deviation between XCO2,cal and
XCO2,exp was determined and , and were further varied until the pCH4,in = pH2O,in = 0 bar, mod,CO2 = 1.6–2.7 kg·s/mol (open symbols) and
deviation reached a minimum. mod,CO2 = 8–13 kg·s/mol (closed symbols)).

900
J. Lefebvre et al. Fuel 239 (2019) 896–904

250 partial pressures leads only to a further reduction in r̄2PM of about 10%.
pH ,in = 1 bar This trend has already been observed by Lim et al. [20]. A strong ad-
CO2 reaction rate r2PM / mmol/(kg·s)

sorption of H2O on the catalyst active sites preventing adsorption of


2

pH ,in = 2 bar
200
2
reactants can explain the effect of H2O on the CO2 reaction rates. The
pH ,in = 3 bar
2 oxidation of catalyst active sites with increasing pH2O, as e.g. reported in
pH ,in = 4 bar
2
Fischer-Tropsch synthesis [40], represents another explanation. Tem-
pH ,in = 5 bar perature programed experiments as well as spectroscopic investigations
150 may help clarifying this phenomenon. However, this is out of the scope
2

of this kinetic study.


Experimental data shown in Fig. 7 were obtained for a H2/CO2 ratio
of 4. In order to see the influence of H2O for different H2/CO2 ratios
100
further experiments were carried out. The results of the corresponding
investigations are shown in Fig. 8, where the inlet H2O partial pressure
is varied from 0 to 0.8 bar for H2/CO2 ratios ranging from 3 to 5 at a
50 temperature of 280 °C. Similar trends can be observed for all in-
0,25 0,75 1,25 1,75 vestigated H2/CO2 ratios. Thus, contrary to the CO2 influence, the H2O
CO2 partial pressure pCO ,in / bar effect on the CO2 reaction rates does not depend on the H2/CO2 ratio.
Altogether, the experimental H2O reaction order derived from loga-
2

Fig. 5. Influence of inlet CO2 partial pressure on the CO2 reaction rate for inlet rithmic linearization does not vary with temperature or H2/CO2 ratio
H2 partial pressures between 1 and 5 bar ( pR = 9.2 bar, T = 260 °C, and is about 0.1. Considering that the negative influence of pH2O,in on
pCH4,in = pH2O,in = 0 bar, mod,CO2 = 1.6–2.7 kg·s/mol, closed symbols: H2/ the CO2 methanation reaction kinetics is due to an adsorption effect, the
CO2 ≥ 4, open symbols: H2/CO2 < 4). low variation of H2O reaction order with temperature is characteristic
of a small adsorption enthalpy.
In Fig. 4, the influence of inlet CO2 partial pressure pCO2,in on the
CO2 reaction rate r̄2PM is shown for an inlet H2 partial pressure of 4 bar.
3.2. Reaction rate equation
An increase in pCO2,in has a little positive effect on r̄2PM . At 300 °C, r̄2PM is
increased by about 17% when pCO2,in rises from 0.75 to 1.25 bar. This
Using the experimental results described in Section 3.1 excluding
trend is more significant for higher temperatures. Accordingly, the CO2
the experiments with H2O in the reactor feed, a power law kinetic rate
reaction order derived from logarithmic linearization of the experi-
equation has been derived from least-squares minimization. The re-
mental data shown in Fig. 4 rises from 0.07 to 0.3 from 200 to 300 °C. In
sulting rate equation is shown in Eq. (14).
literature, a higher positive influence of increasing pCO2 on the CO2
reaction rate has been reported, and most of the published reaction rate 79378 cH0.42 · cCO
0.1
2
equations for CO2 methanation use a CO2 reaction order between 0.3 r2PM = 4.54469·105·exp · ·K
R ·T (1 + 1·cH2O)0.1 (14)
and 1 [21–27].
In Fig. 4, the influence of pCO2,in on the CO2 reaction rate is shown An activation energy EA of 79 kJ/mol – typical for CO2 methanation
for near stoichiometric H2/CO2 ratios of 3–5, typical of two-phase – is retrieved. The parity plot between the experimental CO2 conversion
methanation conditions. However, the findings of Fig. 4 may not be XCO2,exp and the CO2 conversion XCO2,cal calculated with Eq. (14) is il-
relevant for a technical three-phase methanation process, as H2/CO2 lustrated in Fig. 9. A good agreement of experimental results and model
ratios between 1 and 2 are typical for three-phase CO2 methanation is obtained. A standard deviation between XCO2,exp and XCO2,cal of 7.6%
conditions at the catalyst surface. The effect of pCO2,in on the CO2 re- is achieved with the rate described in Eq. (14), assessing a normal
action rate for a H2/CO2 ratio between 0.8 and 6.6 (i.e. pH2,in between 1 distribution.
and 5 bar) and a temperature of 260 °C is shown in Fig. 5. For Experiments with H2O in the reactor feed (grey-marked areas in
pH2,in < 4 bar, i.e. sub-stoichiometric conditions when pCO2,in = 1 bar, Fig. 9) cannot be modeled properly with the rate equation described in
an increase in pCO2,in leads to a small increase in r̄2PM , while the increase
in r̄2PM is more significant for pH2,in 4 bar. The CO2 reaction order 250
derived from these experiments is 0.07 for sub-stoichiometric H2/CO2 200 °C
ratios (i.e. three-phase methanation conditions), and 0.13 for H2/CO2 220 °C
CO2 reaction rate r2PM / mmol/(kg·s)

ratios 4 (i.e. for two-phase methanation conditions). According to our 200 240 °C
knowledge, this observation has never been reported in the literature, 260 °C
because CO2 methanation is usually investigated for stoichiometric H2/ 280 °C
300 °C
CO2 ratios. This effect has been proven by reproduced experiments for 150
both investigated catalyst samples and for different space time velo-
cities. The reduced influence of pCO2,in on r̄2PM for substoichiometric H2/
CO2 ratios can be explained by the lack of adsorbed H2 on the catalyst 100
surface relative to adsorbed carbon species.
The influence of inlet CH4 partial pressure on the CO2 reaction rate
is shown in Fig. 6 for pCO2,in = 1 bar, a H2/CO2 ratio of 4 and tem- 50
peratures ranging from 200 to 300 °C. As expected, the CO2 reaction
rate is almost insensitive to pCH4,in at any investigated temperature,
which is in agreement with most literature [19,20]. 0
0,0 0,4 0,8 1,2
In Fig. 7, the effect of inlet H2O partial pressure pH2O,in on r̄2PM is
depicted for temperatures ranging from 240 to 300 °C, a CO2 partial CH4 partial pressure pCH ,in / bar
4

pressure of 1 bar and a H2/CO2 ratio of 4. At 300 °C, a H2O partial


Fig. 6. Influence of inlet CH4 partial pressure on the CO2 reaction rate for
pressure of 0.4 bar at the reactor inlet results in a strong decrease in r̄2PM
temperatures between 200 and 300 °C ( pR = 9.2 bar, pCO2,in = 1 bar,
of about 30%. This decrease is less pronounced with decreasing tem-
pH2,in = 4 bar, pH2O,in = 0 bar, mod,CO2 = 2 kg·s/mol (open symbols) and
perature. However, a further increase in pH2O,in from 0.4 bar to higher
mod,CO2 = 10 kg·s/mol (closed symbols)).

901
J. Lefebvre et al. Fuel 239 (2019) 896–904

250 Eq. (14). The calculated CO2 conversion rates are systematically 30%
240 °C higher than the experimental CO2 conversions. This corresponds to the
260 °C
CO2 reaction rate r2PM / mmol/(kg·s)

effect described in Section 3.1: addition of water to the reactor feed


200 280 °C
drastically reduces r2PM by about 30%. As the rate equation described in
300 °C
Eq. (14) cannot properly describe the H2O experiments, another rate
equation has been derived for these experiments with least-square
150
minimization. It is expressed in Eq. (15). This new rate equation can
predict CO2 reaction rates with a standard deviation of 13.3%. The only
100
difference between Eqs. (14) and (15) is the pre-exponential factor k 0 .
Eq. (15) is also very similar to the kinetic rate equation developed for

50 Table 2
Measurement uncertainties.

0 Parameter T ci *
0,0 0,4 0,8 1,2 1,6 2,0
Uncertainty ±2 K ±4%
H2O partial pressure pH O,in / bar
* 3% uncertainties are related to the pressure sensor accuracy and 1% to
2

Fig. 7. Influence of inlet H2O partial pressure on the CO2 reaction rate for the GC accuracy.
temperatures between 240 and 300 °C ( pR = 9.2 bar, pCO2,in = 1 bar,
pH2,in = 4 bar, pCH4,in = 0 bar, mod,CO2 = 2 kg·s/mol).
10

Change in CO2 reaction rate ∆r2PM / %


120
(H2/CO2)in = 3
CO2 reaction rate r2PM / mmol/(kg·s)

(H2/CO2)in = 4 5

90
(H2/CO2)in = 5

60

-5

30

-10
cH ± 4 % cCO ± 4 % cH O ± 4 % T ± 2 K Extreme case
0 2 2 2

0,0 0,4 0,8 1,2 Parameter


H2O partial pressure pH O,in / bar
2 Fig. 10. Sensitivity analysis on the kinetic rate equation given in Eq. (14), valid
in absence of H2O in the reactor feed. The extreme case is obtained by setting
Fig. 8. Influence of inlet H2O partial pressure on the CO2 reaction rate for H2/
simultaneously all concentration and temperature uncertainties to their max-
CO2 ratios between 3 and 5 ( pR = 9.2 bar, T = 280 °C, pCO2,in = 1 bar,
imum or minimum value.
pCH4,in = 0 bar, mod,CO2 = 2 kg·s/mol).

0,8 120
Calculated CO2 reaction rate r3PM,cal / mmol/(kg·s)

200 °C
220 °C 230 °C
Calculated CO2 conversion XCO ,cal / -

220 °C +10 %
100
240 °C 250 °C + 10%
240 °C
260 °C 270 °C
2

0,6 260 °C 280 °C 290 °C - 10%


280 °C -10 % 80 300 °C 310 °C
300 °C 320 °C
0,4 60

40
0,2

20

0,0 0
0,0 0,2 0,4 0,6 0,8 0 30 60 90 120
Experimental CO2 conversion XCO ,exp / - Experimental CO2 reaction rate r3PM,exp / mmol/(kg·s)
2

Fig. 9. Parity plot between experimental and calculated CO2 conversions for Fig. 11. Parity plot between CO2 reaction rates measured in a three-phase
2PM experiments. Calculated CO2 conversions using Eq. (14). Grey-marked system and CO2 reaction rates calculated with a two-phase kinetic rate equa-
areas represent the experiments with H2O in the reactor feed. tion. Calculated reaction rates are determined with Eq. (15).

902
J. Lefebvre et al. Fuel 239 (2019) 896–904

three-phase CO2 methanation (see Eq. (2)). A detailed comparison be- and gas concentrations in the liquid phase from three-phase experi-
tween two-phase and three-phase methanation kinetics is carried out in ments are implemented in Eq. (15). The rate equation described in Eq.
the following section. (15) is preferred to the rate equation expressed in Eq. (14), as it takes
into account the presence of H2O in the reactor feed. Indeed, three-
79378 cH0.4 ·c 0.1
r2PM,H2O = 3.24621·105·exp · 2 CO2
·K phase methanation experiments were conducted in a CSTR. As such,
R ·T (1 + 1·cH2O)0.1 (15) H2O is always present in the reaction system. The results of this study
Another type of kinetic rate equation, e.g. a Langmuir-Hinshelwood are shown in a parity plot illustrated in Fig. 11.
type, might solve the issue related to H2O but the simplicity of Eq. (9) is Fig. 11 shows that a very good agreement between 91 experimental
preferred, as only two sets of kinetic parameters depending on the re- three-phase CO2 reaction rates, r3PM,exp , and CO2 reaction rates calcu-
actor feed composition, dry or wet, are sufficient to describe the ex- lated with two-phase methanation rate equation, r3PM,cal , is obtained.
perimental data over a broad parameter range. Assuming a normal distribution, a standard deviation between experi-
To understand the discrepancies between calculated and experi- mental and calculated reaction rates of 5.3% is reached. As a two-phase
mental reaction rates represented in Fig. 9, a sensitivity analysis on the kinetic rate equation is able to estimate three-phase methanation ex-
reaction rate equation given in Eq. (14) was carried out. For this ana- periments, it is confirmed that the liquid phase employed in three-phase
lysis, the reaction temperature and the CO2, H2 and H2O concentrations methanation has no relevant influence on the CO2 methanation ki-
were varied according to the corresponding uncertainties listed in netics. Fig. 11 also confirms that gas concentrations in the liquid phase
Table 2. An extreme case scenario was obtained by setting simulta- and not gas partial pressures in the gas phase are the relevant para-
neously the uncertainty of the parameters to their maximum or meters to describe the three-phase CO2 methanation reaction kinetics.
minimum value.
Fig. 10 shows the influence of measurement uncertainties on the
4. Summary
true CO2 reaction rate r2PM. Temperature has the strongest impact on
r2PM followed by cH2 , cCO2 , and cH2O . The decreasing influence of gas
The objective of the present study was to validate the postulates of a
concentration from H2 to H2O is directly related to the gas species re-
previous publication [13] that a liquid phase does not influence the
action order illustrated in Eq. (14), while the temperature impact is
intrinsic reaction rate of CO2 methanation. For this purpose, CO2 me-
related to the activation energy of reaction. Together, the measurement thanation experiments were carried out using a plug flow laboratory
uncertainties can lead to a deviation in r2PM of about 9%. These un-
fixed-bed reactor, i.e. a reaction system without liquid phase.
certainties can therefore explain the standard deviation between ex- Using the results of 213 validated experiments, a power law kinetic
perimental and calculated XCO2 observed in Fig. 9.
rate equation has been developed, which describes two-phase metha-
nation kinetics on a commercially available catalyst for inlet CO2 par-
3.3. Comparison of two-phase and three-phase methanation kinetics
tial pressures of 1 bar and a temperature range from 200 to 300 °C.
The two-phase methanation kinetic rate equation can describe
To verify the postulates of the previous publication [13], two-phase
three-phase methanation experiments with good agreement, i.e. an
and three-phase CO2 reaction rates are compared to each other: similar
additional liquid phase does not influence the intrinsic reaction rate but
reactions rates should be obtained at similar operating conditions in
the concentration of reacting species on the catalyst surface and gas
both reaction systems if there is no liquid phase influence on CO2 me-
concentration, not gas partial pressure, is the relevant parameter to
thanation reaction kinetics and if gas concentration in the liquid phase
describe the CO2 methanation reaction kinetics.
is the relevant kinetic parameter to describe CO2 methanation reaction
kinetics. It is not possible to compare directly the CO2 reaction rates
measured in two-phase and three-phase systems, as different types of Acknowledgements
reactors were used for the experiments (PFR and CSTR, respectively).
However, it is possible to compare CO2 reaction rates calculated with This work was supported by Bundesministerium für Wirtschaft und
kinetic rate equations. In order to do this, the 2PM kinetic rate equation Energie, Germany under the scope of RegEnKibo project
expressed in Eq. (15) is used to calculate 3PM reaction rates from ex- (Förderkennzeichen 03ET7542F). Special thanks go to Ulli Hammann
periments gathered in the previous publication [13], i.e. temperatures for his assiduous experimental work.

Appendix 1

Technical drawing

Heating zone I Heating zone II Heating zone III

Gas inlet Gas outlet


Fig. 12. Schematic illustration of the lab-scale fixed-bed reactor operated in vertical position. Scheme is adapted from [41].

903
J. Lefebvre et al. Fuel 239 (2019) 896–904

Appendix 2

Correlations and material properties

The axial Peclet number Peax can be calculated using Eq. (16) [45].

1 1 1 bed 1
= +
Peax Pe 2 (16)
The Peclet number Pe is defined as following (Eq. (17)) [45]:
uG· dP
Pe =
D1,2 (17)
2
The binary diffusion coefficient D1,2 (in cm /s) is estimated with Eq. (18) [52]:
0.00143·T1.75 (M1 1 + M2 1 )1/2)
D1,2 =
p 2 ( 1/3 1/3 2
v1 + v 2 ) (18)
In this publication, subscript 1 corresponds to CO2 and subscript 2 to H2. Mi is the molar mass in g/mol, and vi is the diffusion volume. For H2,
the diffusion volume is 6.12, and 26.9 for CO2 according to [52].

[27] Dalmon JA, Martin GA. The kinetics and mechanism of carbon monoxide methanation
References
over silica-supported nickel catalysts. J Catal 1983;84:45–54.
[28] Chiang JH, Hopper JR. Kinetics of the hydrogenation of carbon dioxide over supported
[1] Bundesministerium für Wirtschaft und Energie (BMWi), Fünfter Monitoring-Bericht zur nickel. Ind Eng Chem Prod Res Dev 1983;22:225–8.
Energiewende, http://www.bmwi.de/Redaktion/DE/Publikationen/Energie/fuenfter- [29] Dew JN, White RR, Sliepcevich CM. Hydrogenation of carbon dioxide on nickel-kie-
monitoring-bericht-energie-der-zukunft.html, 2016 (accessed 25.01.2018). selguhr catalyst. Ind Eng Chem 1955;47:140–6.
[2] Bailera M, Lisbona P, Romeo LM, Espatolero S. Power to Gas projects review: lab, pilot [30] Meng F, Li Z, Liu J, Cui X, Zheng H. Effect of promoter Ce on the structure and catalytic
and demo plants for storing renewable energy and CO2. Renew Sustain Energy Rev performance of Ni/Al2O3 catalyst for CO methanation in slurry-bed reactor. J Nat Gas Sci
2017;69:292–312. Eng 2015;23:250–8.
[3] Götz M, Lefebvre J, Mörs F, McDaniel Koch A, Graf F, Bajohr S, et al. Renew Energy [31] Meng F, Li Z, Ji F, Li M. Effect of ZrO2 on catalyst structure and catalytic methanation
2016;85:1371–90. performance over Ni-based catalyst in slurry-bed reactor. Int J Hydrogen Energy
[4] Sterner M, Stadler I. Energiespeicher – Bedarf, Technologien Integration. Berlin: Springer 2015;40:8833–43.
Vieweg; 2014. [32] Gao Y, Meng F, Ji K, Song Y, Li Z. Slurry phase methanation of carbon monoxide over
[5] Hashimoto K, Kumagai N, Izumiya K, Takano H, Kato Z. The production of renewable nanosized Ni–Al2O3 catalysts prepared by microwave-assisted solution combustion. Appl
energy in the form of methane using electrolytic hydrogen generation. Energy Sustain Soc Catal A 2016;510:74–83.
2014;4:17. [33] Gao Y, Meng F, Li X, Wen JZ, Li Z. Factors controlling nanosized Ni-Al2O3 catalysts
[6] Kopyscinski J, Schildhauer TJ, Biollaz SMA. Production of synthetic natural gas (SNG) synthesized by solution combustion for slurry-phase CO methanation: the ratio of redu-
from coal and dry biomass – a technology review from to 2009. Fuel cing valences to oxidizing valences in redox systems. Catal Sci Technol 2016.
1950;89(2010):1763–83. [34] Meng F, Song Y, Li X, Cheng Y, Li Z. Catalytic methanation performance in a low-tem-
[7] Rönsch S, Schneider J, Matthischke S, Schlüter M, Götz M, Lefebvre J, et al. Review on perature slurry-bed reactor over Ni–ZrO2 catalyst: effect of the preparation method. J Sol-
methanation – From fundamentals to current projects. Fuel 2016;166:276–96. Gel Sci Technol 2016:1–10.
[8] Rönsch S, Matthischke S, Müller M, Eichler P. Dynamische Simulation von Reaktoren zur [35] Meng F, Li X, Li M, Cui X, Li Z. Catalytic performance of CO methanation over La-pro-
Festbettmethanisierung. Chem Ing Tech 2014;86:1198–204. moted Ni/Al2O3 catalyst in a slurry-bed reactor. Catal React React Eng EuropaCat V
[9] Lefebvre J, Götz M, Bajohr S, Reimert R, Kolb T. Improvement of three-phase methanation Limerick 2017;2001(313):1548–55.
reactor performance for steady-state and transient operation. Fuel Process Technol [36] Ji K, Meng F, Li Z. Ni-based catalysts prepared by impregnation combustion method for
2015;132:83–90. CO methanation in a slurry-bed reactor. Asia-Pac J Chem Eng 2016;11:151–7.
[10] Götz M, Methanisierung im Dreiphasen-Reaktor. Dissertation, Karlsruhe; 2014. [37] Gao Y, Meng F, Cheng Y, Li Z. Influence of fuel additives in the urea-nitrates solution
[11] Deen NG, Mudde RF, Kuipers JAM, Zehner P, Kraume M. Bubble columns. Ullmann's combustion synthesis of Ni-Al2O3 catalyst for slurry phase CO methanation. Appl Catal A
encyclopedia of industrial chemistry. Weinheim, Germany: Wiley-VCH Verlag GmbH & 2017;534:12–21.
Co. KGaA; 2000. [38] Li X, Meng F, Cheng Y, Gao Y, Li Z. Catalytic methanation in a slurry-bed reactor over Ni/
[12] Nedeltchev S, Schumpe A. Slurry reactors. Handbook of heterogeneous catalysis. Wiley- SiO2 catalysts. React Kinet Mech Catal 2017;122:525–38.
VCH Verlag GmbH & Co. KGaA; 2008. [39] Zhang J, Bai Y, Zhang Q, Wang X, Zhang T, Tan Y, et al. Low-temperature methanation of
[13] Lefebvre J, Trudel N, Bajohr S, Kolb T. A study on three-phase CO2 methanation reaction syngas in slurry phase over Zr-doped Ni/Al2O3 catalysts prepared using different
kinetics in a continuous stirred-tank slurry reactor. Fuel 2018;217:151–9. methods. Fuel 2014;132:211–8.
[14] Götz M, Ortloff F, Reimert R, Basha O, Morsi BI, Kolb T. Evaluation of organic and ionic [40] Iglesias GM, de Vries C, Claeys M, Schaub G. Chemical energy storage in gaseous hy-
liquids for three-phase methanation and biogas purification processes. Energy Fuels drocarbons via iron Fischer-Tropsch synthesis from H2/CO2—Kinetics, selectivity and
2013;27:4705–16. process considerations. Catal Today 2015:184–92.
[15] Graaf GH, Stamhuis EJ, Beenackers AACM. Kinetics of low-pressure methanol synthesis. [41] Iglesias González M. Gaseous hydrocarbon synfuels from H2/CO2 based on renewable
Chem Eng Sci 1988;43:3185–95. electricity: kinetics, selectivity and fundamentals of fixed-bed reactor design for flexible
[16] Graaf GH, Winkelman JGM, Stamhuis EJ, Beenackers AACM. Kinetics of the three phase operation. PhD Thesis, Karlsruhe, Karlsruhe; 2015.
methanol synthesis. Chem Eng Sci 1988;43:2161–8. [42] Weisz PB, Prater CD. Interpretation of measurements in experimental catalysis. Adv Catal
[17] Graaf GH, Sijtsema PJJM, Stamhuis EJ, Joosten GEH. Chemical equilibria in methanol 1954:143–96.
synthesis. Chem Eng Sci 1986;41:2883–90. [43] Anderson JBA. Criterion for isothermal behaviour of a catalyst pellet. Chem Eng Sci
[18] Graaf GH, Smit HJ, Stamhuis EJ, Beenackers Anthonius ACM. Gas-liquid solubilities of 1963;147.
the methanol synthesis components in various solvents. J Chem Eng Data [44] Mears DE. Tests for transport limitations in experimental catalytic reactors. Ind Eng Chem
1992;37:146–58. Proc Des Dev 1971;10:541–7.
[19] Koschany F, Schlereth D, Hinrichsen O. On the kinetics of the methanation of carbon [45] Tsotsas E. M7 Wärmeleitung und Dispersion in durchströmten Schüttungen. VDI-
dioxide on coprecipitated NiAl(O)x. Appl Catal B 2016;181:504–16. Wärmeatlas. Berlin, Heidelberg: Springer Berlin Heidelberg; 2013. p. 1517–34.
[20] Yang Lim J, McGregor J, Sederman AJ, Dennis JS. Kinetic studies of CO2 methanation [46] Binder GG, White RR. Synthesis of methane from carbon dioxide and hydrogen. Chem
over a Ni/Al2O3 catalyst using a batch reactor. Chem Eng Sci 2016;141:28–45. Eng Process 1950;46:563–74.
[21] Weatherbee GD, Bartholomew CH. Hydrogenation of CO2 on group VIII metals. J Catal [47] Kaltenmaier K. Untersuchungen zur Kinetik der Methanisierung von CO2-reichen Gasen
1982;77:460–72. bei höheren Drücken Dissertation, Karlsruhe 1988.
[22] Ibraeva ZA, Nekrasov NV, Gudkov BS, Yakerson VI, Beisembaeva ZT, Golosman EZ, et al. [48] Xu J, Froment GF. Methane steam reforming, methanation and water-gas shift: I. Intrinsic
Kinetics of methanation of carbon dioxide on a nickel catalyst. Theor Exp Chem kinetics. AIChE J 1989;35:88–96.
1991;26:584–8. [49] Hubble RA, Lim JY, Dennis JS. Kinetic studies of CO2 methanation over a Ni/[gamma]-
[23] Kai T, Takahashi T, Furusaki S. Kinetics of the methanation of carbon dioxide over a Al2O3 catalyst. Faraday Disc 2016;192:529–44.
supported Ni-La2O3 catalyst. Can J Chem Eng 1988;66:343–7. [50] Mallard WG, Linstrom PJ. NIST chemistry webbook. NIST Standard Ref Database
[24] Inoue H, Funakoshi M. Kinetics of methanation of carbon monoxide and carbon dioxide. J 1998;69.
Chem Eng Jpn 1984;17:602–10. [51] Kapteijn F, Moulijn JA. Laboratory catalytic reactors. Handbook of heterogeneous cata-
[25] Šolc M. Kinetik der hydrierung des kohlendioxyds zu methan an einem nickel-chrom(III)- lysis. Wiley-VCH Verlag GmbH & Co. KGaA; 2008.
oxyd-katalysator. Collect Czech Chem Commun 1962;27:2621–7. [52] Kleiber M, Joh R. Da1 Berechnungsmethoden für Stoffeigenschaften. VDI-Wärmeatlas.
[26] van Herwijnen T, van Doesburg H, De Jong WA. Kinetics of the methanation of CO and Berlin, Heidelberg: Springer Berlin Heidelberg; 2006. p. 103–32.
CO2 on a nickel catalyst. J Catal 1973;28:391–402.

904

Você também pode gostar