You are on page 1of 9

Journal of CO₂ Utilization 25 (2018) 128–136

Contents lists available at ScienceDirect

Journal of CO2 Utilization


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

Intrinsic kinetics of CO2 methanation over an industrial nickel-based T


catalyst

C.V. Miguel, A. Mendes, L.M. Madeira
LEPABE, Chemical Engineering Department, Faculty of Engineering, University of Porto, Rua Dr. Roberto Frias s/n, 4200-465 Porto, Portugal

A R T I C LE I N FO A B S T R A C T

Keywords: The intrinsic kinetics of CO2 methanation over an industrial nickel-based catalyst was determined for a tem-
Carbon dioxide perature range between 250 °C to 350 °C. The kinetic experiments were performed operating an integral fixed-
Methanation bed reactor far from equilibrium conditions, in the absence of heat and mass resistances and at the atmospheric
SNG pressure. Five mechanistic-based models for describing the reaction kinetics were taken from literature and used
Kinetics
for fitting the experimental reaction rates. Model discrimination was based on the assessment of the thermo-
Nickel catalyst
Sabatier reaction
dynamic consistency and statistical significance of inherent parameters (for 95% confidence level). Comparison
of the adequacy of fit between accepted models was done through the determination of the corresponding F-
values to select the best model. The selected model assumes a formyl intermediate mechanism with a hydroxyl
group being the most abundant species. The proposed reaction kinetics was further validated by the simulation
of an isothermal plug-flow reactor operating at the same experimental conditions employed in this work, where
a reasonable agreement between model predictions and the observed values was obtained.

1. Introduction Carbon dioxide methanation reaction (Eq. (1)) is thermo-


dynamically favoured at low temperatures and high pressures [5]. The
The main application for the methanation reaction (also known as equilibrium constant (Keq) dependence with the absolute temperature
the Sabatier reaction) has been since long ago the removal of carbon (T) can be retrieved from the equation provided by Lunde and Kester
oxides traces from hydrogen-rich feed streams in ammonia plants [1,2] [6]:
(Eqs. (1) and (2)). It has also been proposed to produce synthetic nat- 1.0
K eq = e⎣ ( 1.987 )
⎡ (56000 T 2+ 34633 T -16.4ln(T ) +0.00557T ) +33.165⎤
ural gas (SNG), particularly in the 60’s decade, because of augmented ⎦ (3)
natural gas demand. At that time, SNG production from coal was en-
The heat released by the reaction is the major difficulty to handle at
visioned as an alternative pathway to assure security of supply in any
industrial scale [3]; the adiabatic rise in temperature per each percent
event of natural gas shortage and research on the topic was strongly
of CO2 converted is 60 °C [7]. Hence, methanation usually takes place
financed, particularly by the USA [3].
in a series of adiabatic fixed-bed reactors with inter-bed gas recycling
CO2 +4H2 ⇌ CH 4 +2H2 O ΔHr298 K = - 165 kJ⋅mol-1 (1) cooling or in fluidized-bed reactors [3,8]. Structured catalysts such as
metal coated foams have also been envisaged for this reaction due to
CO+3H2 ⇌ CH 4 + H2 O ΔHr298 K = - 206 kJ⋅mol-1 (2) their improved heat transfer capacity (e.g. [9,10]).
Nowadays, CO2 methanation reaction has gained a renewed interest Ni-based catalysts supported on various solids (e.g. Al2O3, SiO2,
in the scope of Power-to-Gas applications (PtG), a concept where sur- CeO2, etc.) are the most studied and commercialized catalysts at high
plus renewable electricity is transformed into hydrogen (via H2O elec- temperatures (i.e. > 250 °C), a range where the formation of dangerous
trolysis) and afterwards into methane. This last step makes the process nickel carbonyl (Ni(CO)4) is avoided, while more expensive Ru-based
more flexible since methane can be more easily stored and transported catalysts are best options at low temperatures (< 200 °C) [7,11]. Gen-
than hydrogen, enabling the integration and balance of the power grid erally, catalysts which are effective for CO methanation (Eq. (2)) are
with the gas grid. Methane wide range of end-use possibilities (e.g. also effective for CO2 methanation, at least for streams having low COx
vehicle fuel, for heat and power production, intermediate to obtain concentrations, as found in hydrogen purification processes [11].
other chemicals) contributes to process flexibility and versatility [4]. Table 1 lists main manufacturers and characteristics of some


Corresponding author.
E-mail address: mmadeira@fe.up.pt (L.M. Madeira).

https://doi.org/10.1016/j.jcou.2018.03.011
Received 2 January 2018; Received in revised form 6 March 2018; Accepted 12 March 2018
2212-9820/ © 2018 Elsevier Ltd. All rights reserved.
C.V. Miguel et al. Journal of CO₂ Utilization 25 (2018) 128–136

Nomenclature R Ideal gas constant (8.314 J mol−1 K−1)


T Absolute temperature (K)
a′ Specific external area of the catalyst particle (a' = 6/dp for Tb Temperature in the bulk phase (K)
spheres) (m−1) Ts Temperature at the catalyst particle surface (K)
Ca Carberry number (−) Tm Mean temperature of range (Tm = 300 °C) (K)
Cb CO2 concentration in the bulk phase (mol m−3) uo Fluid superficial velocity (m s−1)
Cs CO2 concentration at the surface of the catalyst particle Vr Reactor volume (m3)
(mol m−3) Wcat Catalyst weight (kg)
dp Particle diameter (m) XCO2 CO2 conversion (−)
Deff Effective mass diffusivity in the catalyst (m2 s−1) yi Mole fraction of species i (−)
Dr Reactor diameter (m) YCH4 Methane yield (−)
Ea Activation energy (J mol−1) z Dimensionless reactor length (−)
out
FCH 4
CH4 outlet molar flow rate (mol s−1)
in
FCO 2
CO2 inlet molar flow rate (mol s−1) Greek letters
h Heat transport coefficient between gas and particle
(W m−2 K−1) ΔH Enthalpy of adsorption (J mol−1)
k Reaction rate constant (mol g−1 s−1 Pa−n) ΔHr Heat of reaction (J mol−1)
kg Mass transport coefficient (m s−1) βe Dimensionless number for extraparticle heat transport (−)
k0 Pre-exponential factor of rate constant (mol g−1 s−1 Pa−n) βi Dimensionless number for intraparticle heat transport (−)
K0 Pre-exponential factor of adsorption equilibrium constant γe External Arrhenius number (−)
(Pa−1 or Pa−0.5) γi Internal Arrhenius number (−)
K Adsorption equilibrium constant (Pa−1 or Pa−0.5) ε Column porosity (0.36–0.38) (−)
Keq Equilibrium constant for methanation (Pa−2) ρg Density of the gas mixture (kg m−3)
L Reactor length (m) ρp Density of the catalyst particle (g m−3)
Lp Characteristic catalyst dimension (Lp = dp/6 for spheres) η Internal effectiveness factor (−)
= (m) ϕ Generalized Thiele modulus (−)
Mi Molar weight of species i (kg mol−1) ηϕ2 Wheeler Weisz modulus (−)
nr Reaction order (−) λg Thermal conductivity of the gas mixture (W m−1 K−1)
pi Partial pressure of species i (Pa) λ eff Effective thermal conductivity of the catalyst particle
P Total pressure (Pa) (W m−1 K−1)
Q Volumetric feed flow rate (mLn min−1) μi Viscosity of pure species i (Pa s)
rCH4 Reaction rate of methane formation (mol g−1 s−1) μg Viscosity of the gas mixture (Pa s)
robs Observed reaction rate (mol g−1 s−1) νi Stoichiometric coefficient of species i (−)
rmod Model predicted reaction rate (mol g−1 s−1) Ψ Inert dilution factor (−)

industrial methanation catalysts. The catalysts are available in several available commercial catalysts and their properties, useful information
shapes and some are supplied in a pre-reduced form (e.g. Katalco 11-4R about their preparation for use, handling instructions and safety pre-
or PK-7R), which simplifies process start-up. Industrial methanation cautions, besides operation issues of some industrial processes where
catalysts lifetime ranges from 5 to 10 years, although some manu- methanation intervenes.
facturers report a period up to 24 years [1,7]. Common poisons are This work determines the intrinsic reaction kinetics over an in-
sulphur and arsenic compounds, particularly for nickel catalysts [2]. dustrial nickel-based catalyst. Knowing the reaction kinetics is funda-
Industrial catalysts for hydrogenation of carbon oxides has been the mental for modelling, simulation and optimization of conventional or
subject of a detailed review by Golosman and Efremov [7], covering the new reactor concepts (e.g. sorptive reactors [16]). To this end, me-
chanistic-based rate equations available in the literature were con-
sidered and are presented in the following section.
Table 1
List of some industrial methanation catalysts and related characteristics [7,12–15].
2. Mechanisms for CO2 methanation
Commercial Metal Metal P (bar) T (ºC) Shape Size (mm)
reference cont. 2.1. “Carbon intermediate” mechanism
(wt. %)

METH 134 a) Ni 20-25 n.a. > 200 Ball 3.0–6.0 Dalmon and Martin [17] proposed a mechanism for CO and CO2
METH 150 a) Ru n.a. n.a. < 170 Pellet 4.5–4.5 methanation over a Ni/SiO2 catalyst by studying the hydrogenation of
PK-7R b) Ni > 23 n.a. 190–450 Ring 5.0 × 2.5 intermediate species. The authors postulated that in both reactions
Katalco 11-4 c) Ni n.a n.a. n.a. Pellet 5.4 × 3.6
adsorbed carbon monoxide (CO*) is a common intermediate (apart
Katalco 11-4R c) Ni n.a n.a. > 220 Pellet 5.4 × 3.6
Katalco 11-4M c) Ni n.a n.a. n.a. Pellet 3.1 × 3.6 from O* in the latter case). CO* then dissociates into C* and O* while
Katalco 11-4MR c) Ni n.a n.a. > 220 Pellet 3.1 × 3.6 CH4 is produced following C* hydrogenation. Since CO* can be formed
NIAP-07-01 d) Ni 33–39 20–300 180–450 Pellet 5.5 × 4.5 at lower temperatures in the case of CO2 adsorption, a lower activation
NIAP-07-02 d) Ni 32–38 20–300 180–450 Pellet 5.5 × 5.0 energy could explain the higher hydrogenation rate of CO2. The higher
NIAP-07-04 d) Ni 25 n.a n.a Pellet 5.5 × 4.5
TO-2M d) Ni 35–41 1–300 180–450 Pellet 5.5 × 5.0
selectivity of CO2 methanation was tentatively explained by the lower
RKM-3 d) Ru 0.3 300 150 Ball 2.6 - 3.0 C* concentration (because of lower CO2 adsorption capacity) and to the
Manufacturer higher abundance of O*, which may act as a geometric diluent de-
a) Clariant b) Haldor- c) Johnson Matthey d) NIAP Katalizator creasing the C-C formation.
Topsoe
Weatherbee and Bartholomew [18] also used a Ni/SiO2 catalyst and
the mechanism proposed by Dalmon and Martin adequately explained

129
C.V. Miguel et al. Journal of CO₂ Utilization 25 (2018) 128–136

their results. The elementary steps of the mechanism are presented in [23],
Table 2.
0.5
The kinetic rate equation derived, assuming CO* dissociation to C* kpCO p 0.5
2 H2 ⎛ pCH4 pH22O ⎞
rCH4 = 0.5 2 ⎜
1 −
and O* (step 4) as the rate determining step (r.d.s.), is given by Eq. (4), (1 + K H2O pH2O + K H0.5 p 0.5
2 H2
+ Kmix pCO ) pCO2 pH42 K eq ⎟ (7)
2 ⎝ ⎠
0.5
kpCO p 0.5
2 H2 and the term (K H2O pH2O ) stands for the water surface coverage (ΘH2O) .
rCH4 = 2
0.5
⎛1 + K1 pCO2 + K2 p 0.5 p 0.5 + K3 p ⎞
⎜ ⎟
pH0.5 CO2 H2 CO
⎝ 2 ⎠ (4) 3. Experimental
where k is the rate constant, pi is the partial pressure of species i (CO2,
H2, CO) and K1, K2 and K3 are the constants that result from the 3.1. Experimental setup
combination of several adsorption equilibrium and/or reaction rate
constants. An illustration of the experimental setup used in the kinetic tests is
shown in Fig. 1.
Experiments were performed in a stainless steel packed-bed unit
2.2. “Formate intermediate” mechanism
with 12 cm length and 7.75 mm inside diameter, which was placed
inside a tubular oven (model Split from Termolab, Fornos Eléctricos,
Fujita et al. [19] studied the mechanism of CO and CO2 methanation
Lda.) equipped with a 3-zone PID temperature controller (model MR13
over an Ni/Al2O3 catalyst at 180 °C. For CO methanation it was ob-
from Shimaden). The 3 type-K thermocouples used to measure and
served that adsorbed carbon, linear and bridged CO were the pre-
control the oven temperature were placed in contact with the reactor
dominant species, whereas for CO2 methanation the presence of two
wall.
types of adsorbed CO in bridged structures and a formate species
CO2 (99.998%, Air Liquide), H2 (99.9995%, Air Liquide) and N2
(HCOO) were detected. The hydrogenation of adsorbed carbon was
(99.9995% %, Air Liquide) were fed upwardly to the reactor using mass
markedly retarded by the presence of linear CO species. Since the latter
flow controllers (model F201 from Bronkhorst-High Tech). The flow
species has not been observed during CO2 methanation, this could ex-
rate of the stream leaving the reactor was measured with a mass flow
plain the difference between the rate and selectivity of CO and CO2
meter (model F101 from Bronkhorst-High Tech).
methanation reactions [19,20].
Type-K thermocouples allowed recording the temperature histories
Pan et al. [21] recorded the in-situ FTIR spectra of CO2 methanation
in two axial positions of the reactor. The reactor was placed inside the
using a Ni/Al2O3 catalyst at different temperatures (75–375 °C). The
oven so that the temperature in the bed was practically the same along
presence of bidentate formate intermediate was observed for tempera-
its length (maximum difference of 1 °C). The pressure drop was mea-
tures above 225 °C and its hydrogenation was considered the main re-
sured by means of two pressure transducers (model PMP 4010 from
action step. Both Fujita et al. [19] and Pan et al. [21] concluded that
Druck) placed before and after the reactor. The water vapour produced
formate species was adsorbed at the alumina support. A mechanism for
during the course of the reaction was condensed in a home-assembled
CO2 hydrogenation over a commercial nickel-alumina-calcium catalyst
Peltier cold-trap located after the reactor.
(ref.: NKM-4A from NIAP-Katalizator) involving a formate intermediate
The steady-state composition of the (dry) stream leaving the reactor
was given by Ibraeva et al. [22] (cf. Table 3).
was measured using a gas chromatograph (model 1000 from DANI)
In this mechanism, the formate species is obtained through the re-
equipped with a micro-TCD detector (VICI) and a capillary column
action of molecularly adsorbed CO2 and atomic hydrogen (step 3),
(Supelco Carboxen 1010 Plot, 30 m x 0.32 mm i.d. from Sigma
being this the rate determining step of the reaction. The derived kinetic
Aldrich). Hydrogen fraction was obtained through the mass balance.
rate equation is expressed by Eq. (5),
The carbon balance error was lower than 10%.
pH0.5 p
2 CO2
rCH4 = k 0.5
pH2 + KpCO2 (5) 3.2. Kinetic experiments

where K is the adsorption equilibrium constant. Kinetic tests were performed using a commercial nickel catalyst
(METH 134). The catalyst has a NiO content of ca. 20–25 wt. % sup-
2.3. “Formyl intermediate” mechanism ported on calcium aluminate [7,24]. Axial dispersion and gas-solid wall
effects can be neglected when L/dp and Dr/dp ratios are above 50 and
A mechanism comprising a formyl species intermediate has been 10, respectively [25,26], which was guaranteed in the present work.
proposed recently by Koschany et al. [23] for a coprecipitated NiAl(O)x Pressure drop along the bed may be assumed negligible (< 7.4% of
catalyst (cf. Table 4). inlet pressure).
In this mechanism, the formyl species results from the reaction of Temperature was varied between 250–350 °C, a range where com-
adsorbed CO (obtained after CO2 dissociation) and atomic hydrogen. mercial catalysts are designed to operate without activity loss, even
Eq. (6) was derived by Koschany et al. [23] assuming adsorbed CO with occasional temperature excursions up to 650 °C [2]. The activity of
hydrogenation (step 3) as the rate limiting step and CO*, H* and OH* as
the most abundant surface species, Table 2
Carbon intermediate mechanism (from [18]).
0.5
kpCO p 0.5
2 H2 ⎛ pCH4 pH22O ⎞
rCH4 = ⎜1 −
pH2 O
2
pCO2 pH42 K eq ⎟ Elementary step Step
⎛1 + K OH
⎜ + K H0.5 p 0.5 0.5 ⎞ ⎝
+ Kmix pCO2 ⎟ ⎠
0.5
pH 2 H2
⎝ 2 ⎠ (6) H2 + 2* ⇌
CO2 + 2* ⇌
pH2 O
where ⎛K OH 0.5

⎞, (K H0.5 p 0.5 )and (Kmix p 0.5 ) stand for the surface cov-

2 H2 CO2
CO* ⇌
pH CO*+* ⇌
⎝ 2 ⎠
erage of hydroxyl (ΘOH) , atomic hydrogen (ΘH) and carbon monoxide C*+ 4H* ⇌
(ΘCO) , respectively. K eq is the reaction equilibrium constant determined CH 4* ⇌
O*+ H* ⇌
through Eq. (3).
OH*+ H* ⇌
If water is considered as one of the most abundant surface inter- H2 O* ⇌
mediates instead of hydroxyl species, than Eq. (6) is modified as follows

130
C.V. Miguel et al. Journal of CO₂ Utilization 25 (2018) 128–136

Table 3 determined if the experiments are conducted in the absence of both


Formate intermediate mechanism (from [22]). external and internal resistances. In this regard, catalyst pellets were
crushed and sieved to different sizes, namely, between 150–250 μm and
Elementary step Step
250–350 μm; Kiendl et al. [24] used a similar catalyst and reported that
H2 + 2* ⇌ crushing has no influence on the Ni loading. The catalyst particles were
CO2 +* ⇌ then mixed with inert glass beads with the same particle size. Dilution
CO2*+ H* ⇌
of the catalyst with inert glass beads enhances the heat transfer and
HCOO*+ H* ⇌
HCO*+ H* ⇌
minimizes temperature gradients, which was confirmed through direct
CH*+ 3H* ⇌ measurement of the bed temperature. Due to the highly exothermic
OH*+ H* ⇌ nature of the Sabatier reaction (Eq. (1)), the CO2 content in the reactant
mixture was low and balanced in N2 to minimize the temperature in-
crease. An H2/CO2 feed ratio of 4 was chosen for all experiments; at low
Table 4 temperature and atmospheric pressure, H2/CO2 ratios lower than 4
Formyl intermediate mechanism (from [23]).
could lead to low CH4 selectivity and undesired carbon formation, as
Elementary step Step found in a previous thermodynamic study [5].
The mixture of catalyst and inert particles was loaded into the fixed-
H2 + 2* ⇌ bed tubular unit and framed in both ends by means of two stainless steel
CO2 + 2* ⇌
discs (mesh: 10–15 μm). At the bottom of the reactor (i.e. at the inlet) a
CO*+ H* ⇌
CHO*+* ⇌
layer of inert particles allows to even out the flow of the reaction
CH*+ 3H* ⇌ mixture. Catalyst activation consisted in the following protocol:
CH 4* ⇌
O*+ H* ⇌ 1 Purging the entire system with N2 at room temperature;
OH*+ H* ⇌ 2 Heating the packed-bed (1 °C min−1) from room temperature up to
H2 O* ⇌
320 °C at the ambient pressure under a 12.8% H2/N2 stream and
hold 1h at that temperature;
commercial nickel catalysts does not increase significantly above 350 °C 3 Increasing the temperature under N2 flow to the desired value for
[27]. The experiments were performed at the atmospheric pressure. the first kinetic experiment (i.e. 350 °C).
Blank tests using the reactor packed with only glass beads showed the
absence of CO2 conversion in the considered temperature range. A 4-way valve allowed to: 1) keep the reactor under a static N2
The first set of kinetic tests was carried at 350 °C to check the ab- atmosphere while the reactant mixture stream could be prepared and
sence of external mass and heat transport resistances by varying the the flow stabilized and 2) feed the reactant stream to the reactor.
feed flow-rate while the value of the time coordinate Wcat FCO in
was kept Shutdown consisted on flushing the reactor with N2 to remove un-
2
constant (see Sections 4.1.1 and 4.1.2). In the absence of external mass/ reacted carbon oxides and afterwards cooling down the reactor up to
heat transfer effects and in the presence of isothermal operation at room temperature.
350 °C, the same premises are guaranteed at lower temperatures and
lower catalyst activities [25]. 3.3. Computational methods
After assuring the absence of external resistances, the presence of
internal diffusion limitations was assessed using catalyst particles with The parameters of the kinetic equations were estimated through
different sizes (see Section 4.1.2). Intrinsic kinetics can thus be nonlinear regression and statistically tested using the fitnlm function
available in the Matlab software as detailed below (cf. Section 4.2.1).

Fig. 1. Scheme of the experimental setup used in this work. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

131
C.V. Miguel et al. Journal of CO₂ Utilization 25 (2018) 128–136

Nonlinear regression algorithm consisted in the Levenberg-Marquardt different size was nearly the same (Fig. 3).
method (details available in [28]) for the minimization of the sum of
residuals squares. Statistical testing was performed for 95% level of 4.1.3. Calculation of observed reaction rates
significance. out in
Fig. 4 shows the observed methane yield (YCH4 = FCH 4
FCO 2
) as a
The selected kinetic model was validated based on the simulation of function of the ratio between the weight of catalyst (Wcat ) to the CO2
the fixed-bed reactor. The differential equations and respective initial molar feed flow rate (FCO in
2
) at temperatures ranging from 250 °C to
conditions were solved numerically with a fourth-order Runge-Kutta 350 °C.
algorithm also using the Matlab software. The observed reaction rates were obtained using the differential
method [26].
4. Results and discussion
dYCH4
rCH4 =
d(Wcat FCO2 ) (12)
4.1. Kinetic experiments
in
To this end, the relationship between YCH4 and was de-
Wcat FCO 2
4.1.1. Isothermal regime and catalyst stability scribed in terms of polynomial functions for each temperature (see
Reactor layouts were varied to assess the system behaviour re- dashed lines in Fig. 4). The derivative of each polynomial function al-
garding the planned operation conditions (cf. Fig. 2). lowed to extract the observed reaction rates at a given temperature and
in
Isothermal operation must be guaranteed for the determination of Wcat FCO 2
(cf. Eq. (12)). Falsification of the estimated reaction rates due
the intrinsic kinetics. Due to the exothermic nature of CO2 methanation, to a deviation of the experimental conversion caused by catalyst dilu-
the bed temperature rise depends on the number of moles of CO2 pre- tion was discarded based on the correlation proposed by Berger et al.
sent in the feed composition which are converted per unit of time. In [30], where such difference is estimated considering only observable
preliminary experiments, the initial bed temperature increased 8.7 °C parameters. It was found that the average relative deviation between
vs. 2.8 °C by changing the CO2 feed fraction from 12 vol.% to 4 vol. %, the conversion obtained under diluted and undiluted conditions was
respectively, using a reactor with 200 mg of catalyst (reactor R1 in only 1%.
Fig. 2). This effect was not possible to assess for lower CO2 percentages
due to mass flow controllers limitations. In turn, the available heat 4.2. Modelling work
transfer area per mass of catalyst was increased changing the dilution
factor (Ψ) and using a feed composition (vol. %) of 4:16:80 (CO2:H2:N2) 4.2.1. Model discrimination and parameter estimation
in all the subsequent experiments. Isothermal regime was considered A scaling method consisting in temperature centering was adopted
when the variation of the temperature measured in two positions of the for the estimation of rate and adsorption parameters of the proposed
bed length was lower than 1.5 °C after feeding the reactant mixture, kinetic equations, according to Eqs. (13) and (14), respectively.
which was attained for the reactor with configuration R3 in Fig. 2,
under the considered range of total feed flow rates (30–100 mln min−1) ⎡− Ea ⎛ 1 − 1 ⎞ ⎤
k = k 0′ e⎣ R ⎝ T Tm ⎠ ⎦ (13)
and catalyst weights used in the experiments (12–40 mg).
Catalyst stability was checked under reactive conditions for a total ⎡− ΔH ⎛ 1 − 1 ⎞ ⎤
period of approximately 83 h. During this time the catalyst was sub- K = K 0′ e⎣ R ⎝ T Tm ⎠ ⎦ (14)
mitted to reactive conditions in the temperature range of 250–350 °C.
where Tm stands for the mean temperature of the considered range (i.e.
The catalyst was always kept in a nitrogen stream at 350 °C overnight.
573 K) and k 0′ and K 0′ are the constants evaluated at T = Tm.
The observed difference of CH4 yield measured at the beginning and the
Temperature centering reduces the correlation between the pre-ex-
end of this procedure (operation conditions: T = 350 °C, P = Patm,
ponential factor and Ea or ΔH, making the numerical least-squares
in
Wcat FCO = 6.22 gcat h mol−1) was negligible (ca. 0.61%). Therefore, in
2 procedure more robust and less sensitive to poor starting estimates
the kinetic tests it was considered that the catalyst activity remained
[28]. Afterwards, the “true” pre-exponential factors are calculated from
unchanged since they were conducted in less time than the reference
Eqs. (15) and (16), by the Arrhenius and Van’t Hoff equations, re-
period.
spectively.
Ea
4.1.2. Identifying the region of kinetic rate control ⎛ ⎞
k 0 = k 0′ e⎝ RTm ⎠ (15)
The extraparticle and intraparticle mass and heat transport transfer
limitations were assessed based on the criterion listed in Table 5.
The expressions employed for determining the thermodynamic and
transport properties and parameters used throughout Eqs. (8)–(11) are
given in the Supplementary Information (see parameters definitions
and units in the Nomenclature section). The criterion listed in Table 5
were always verified considering the extreme situations, i.e.:

1) Conditions of minimum and maximum reaction rate at 350 °C and;


2) Using either reactor feed or outlet compositions for calculating the
parameters.

The inequalities of Eqs. (8)–(11) were all verified and thus both
extraparticle and intraparticle heat and mass transport limitations
could be neglected. In addition, the effect of external resistances was
also assessed experimentally at 350 °C by determining CO2 conversion
and varying the feed flow rate, while keeping the time coordinate
in
Wcat FCO 2
constant (cf. Fig. 3).
Fig. 3 shows the absence of external resistances for a feed flow rate
range between 30–100 mln min−1. Internal diffusion effects can also be
Fig. 2. Reactor layouts employed to determine the best configuration for the kinetic tests.
neglected since the conversion obtained with catalyst particles of a

132
C.V. Miguel et al. Journal of CO₂ Utilization 25 (2018) 128–136

Table 5
Criteria for negligible transport limitations in steady state kinetic studies [29].

Transport process Criterion Eqs.

External mass transport robs ρ p (8)


Ca = < 0.05
kg a ′ C b
Intraparticle mass transport robs ρ p Lp2 nr + 1 (9)
(Wheeler-Weisz criterion) ηϕ2 = ⋅ < 0.10
Deff Cs 2
External heat transport kg (−ΔHr ) C b Ea (10)
|βe⋅γe⋅Ca| = ⋅ ⋅Ca < 0.05
hT b RT b
Intraparticle heat transport Deff (−ΔHr ) Cs Ea (11)
|βi⋅γi⋅ηϕ2| = ⋅ ⋅ηϕ2 < 0.05
λ eff Ts RTs

Fig. 5. Parity plot of rmod versus robs showing adequacy of fit of proposed model (Eq.
(18)).

minus zero (null hypothesis) divided by the standard deviation of the


parameter. If the obtained t-value is lower than the t-value for α = 0.95
and (k-p) degrees of freedom, or has a large confidence interval in-
cluding zero, it has no significant contribution to the rate equation and
can be deleted [26].
The kinetic models proposed in Section 2 were then rejected based
on the criteria mentioned above. However, the experimental data could
be quite well described after deleting the hydrogen and carbon mon-
oxide coverage terms from the denominator of Eqs. (6) and (7). The
simplified reaction rate equations are as follows:
Fig. 3. Effect of total volumetric feed flow rate on CO2 conversion for constant
Wcat FCO2 = 3.73 g h mol−1 at 350 °C. Dashed line stands for the thermodynamic equili- 0.5
kpCO p 0.5
2 H2 ⎛ pCH4 pH22O ⎞
brium conversion at 350 °C, 1 bar and feed composition.
rCH4 = ⎜1−
2
pCO2 pH42 K eq ⎟
⎛1 + K OH pH2O ⎞ ⎝
⎜ ⎟ ⎠
pH0.5
⎝ 2 ⎠ (18)

0.5
kpCO p 0.5
2 H2 ⎛ pCH4 pH22O ⎞
rCH4 = ⎜1 − p p 4 K eq ⎟
(1 + K H2O pH2O )2 CO2 H2 (19)
⎝ ⎠

The kinetic model assuming water as most abundant species (Eq.


(19)) was however rejected because the obtained F-value was lower
than the F-value for the model expressed by Eq. (18) (209 vs. 259).
Fig. 5 shows the parity plot of the observed reaction rates against
those obtained by the model given by Eq. (18).
Table 6 lists the parameters estimates with a 95% confidence level
(at reference conditions).
Besides the significance of the estimated parameters, the F − value
was also satisfied (i.e. 259 > 2.76). The estimated activation energy
in
Fig. 4. Methane yield as a function of the time coordinate Wcat FCO 2 at various tem-
was 118.7 kJ mol−1, a value close to others reported in literature for
peratures. The continuous line stands for the equilibrium value of YCH4 at 350 °C, 1 bar similar nickel catalysts (e.g. 105.9 kJ mol−1 in [31] and 113.5 kJ mol−1
and feed composition (i.e. the minimum YCH4 for the temperature range considered). in [32]). The adsorption enthalpy of the hydroxyl coverage term is
positive, which is in agreement with the results by Koschany et al. [23].
⎛ ΔH ⎞
The obtained hydroxyl enthalpy of adsorption was 61.6 kJ mol−1, a
K 0 = K 0′ e⎝ RTm ⎠ (16) value which is however 2.75 times higher than the value reported by
Koschany et al. [23].
Model discrimination was based on the thermodynamic consistency
and statistical validity of the estimated parameters and the overall
Table 6
adequacy of fitting given by the F-value with a confidence level α , as
Parameter estimates at reference temperature (573 K) and corresponding lower (LL) and
proposed by Froment and Bischoff [26] (cf. Eq. (17), upper (UL) limits of the 95% confidence interval.
2
∑k rmod
Parameter Units Estimate t-value LL UL
p
F − value = > F (p , k − p ; 1 − α )
∑k (rmod − robs )2
k 0′ (mol⋅g -1⋅h-1) × 100 0.8936 8.959 0.7938 0.9933
k−p (17)
E
R
K 14275 13.34 13205 15345
where k and p stand for the number of experiments and estimated ′
K 0,OH kPa−0.5 0.4326 3.001 0.2885 0.5768
parameters, respectively. The statistical significance of parameters was ΔH
R
K 7414 3.626 5369 9458
assessed through the t-value, which is equal to the parameter estimate

133
C.V. Miguel et al. Journal of CO₂ Utilization 25 (2018) 128–136

3) Negligible mass and heat transport limitations within the catalyst


(internal) and between the catalyst and the bulk gas phase (ex-
ternal);
4) Ideal gas behaviour.
5) Reaction takes place on the catalyst surface.

The model included the estimation of pressure (P) along the di-
mensionless reactor length (z) by the Ergun equation [33]:

dP ⎛ (1 − ε )2μg (1 − ε ) ρg 2⎞
= −L ⎜150 2
uo + 1.75 uo ⎟
dz ⎝
3
ε dp ε 3d p ⎠ (20)

where L is the reactor length, ε is the column porosity, d p is the particle


diameter, μg is the gas mixture viscosity, uo is the fluid superficial ve-
locity and ρg is the gas mixture density. The gas mixture density (ρg ) and
viscosity (μg ) were calculated locally and the equations used for their
estimation are given in the Supplementary Information.
Fig. 6. Model predicted vs. observed values of the methane yield at different tempera- The fluid superficial velocity (uo) was estimated along z by the
tures. following differential equation obtained from the total mass balance to
the reactor in steady-state (Eq. (21)):
4.2.2. Model validation duo u dP W RTL
A one-dimensional pseudo-homogeneous plug-flow model was used =− o − cat ∑ (νi r )
dz P dz Vr P i (21)
to validate the selected kinetic expression against the experimental
data. The following main assumptions have been made: where Wcat is the weight of catalyst, Vr is the reactor volume and νi is the
stoichiometric coefficient of species i (excluding N2).
1) Isothermal operation; The mole fraction of species i along z was estimated by the following
2) Negligible axial and radial dispersion; general equation obtained from the partial mass balance to the reactor

Fig. 7. Simulated values along the dimensionless reactor length of: (a) the species mole fractions and (b) reaction rate at T = 350 °C and for the lower and higher time coordinate values
considered. Red squares stand for the experimental values. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

134
C.V. Miguel et al. Journal of CO₂ Utilization 25 (2018) 128–136

which assumes hydrogen and carbon dioxide dissociation followed by


hydrogenation of adsorbed carbon monoxide to yield a formyl species,
assuming hydroxyl as the most abundant species, showed a good fit to
the experimental data.
An isothermal plug-flow model including the proposed kinetic rate
equation with estimated parameters was used to simulate the fixed-bed
reactor. The reactor model satisfactorily captured the experimental
in
observations in limiting conditions of temperature and Wcat FCO 2
, al-
lowing to validate the proposed (intrinsic) kinetic rate model.
The selected kinetic rate and estimated model parameters can thus
be transferred and used for modeling, simulation and design of in-
dustrial reactors used for CO2 utilization applications, where metha-
nation is standing out among the existing options.

Acknowledgments

This work was the result of the following projects: (i) POCI-01-0145-
Fig. 8. Simulated methane yield profile along the dimensionless reactor length for the
FEDER-006939 (Laboratory for Process Engineering, Environment,
highest and lowest robs conditions measured at 250 °C and 350 °C. Red squares symbols
stand for the experimental values. (For interpretation of the references to colour in this
Biotechnology and Energy–UID/EQU/00511/2013) funded by the
figure legend, the reader is referred to the web version of this article.) European Regional Development Fund (ERDF), through COMPETE2020
– Programa Operacional Competitividade e Internacionalização (POCI)
and by national funds, through FCT - Fundação para a Ciência e a
in steady-state (Eq. (22))
Tecnologia; (ii) NORTE‐01‐0145‐FEDER‐000005 – LEPABE-2-ECO-
dyi y duo W RT L y dP INNOVATION, supported by North Portugal Regional Operational
=− i − (νi r ) cat − i
dz uo dz Vr P uo P dz (22) Programme (NORTE 2020), under the Portugal 2020 Partnership
Agreement, through the European Regional Development Fund (ERDF).
The initial conditions for solving the differential Eqs. (20)– (22) C.V. Miguel is grateful to FCT for his PhD scholarship (SFRH/BD/
were P|z = 0 = Pin , uo |z = 0 = uo0 and yi |z = 0 = yi0 , where the superscript “0” 110580/2015), financed by national funds of the Ministry of Science,
refers to feed conditions. Technology and Higher Education and the European Social Fund (ESF)
Fig. 6 shows the model predictions of methane yield compared to through the Human Capital Operational Programme (POCH).
the observed values. A very good agreement was found between model
predicted and observed values along the large methane yield and Appendix A. Supplementary data
temperature ranges, i.e. from differential up to integral conditions.
Fig. 7 shows the simulation profiles of the species mole fractions and Supplementary material related to this article can be found, in the
methane formation reaction rate along the dimensionless reactor length online version, at doi:https://doi.org/10.1016/j.jcou.2018.03.011.
calculated in maximum and minimum reaction rate conditions at
350 °C. The predicted values at the reactor outlet (i.e. z = 1) are com- References
pared with the experimental values (see the red squares in Fig. 7).
Fig. 7a shows that H2 and CO2 are consumed along all the reactor [1] H.F. Rase, Handbook of Commercial Catalysts: Heterogenous Catalysts, CRC Press,
length while H2O and CH4 are continuously produced because one is Boca Raton, 2000.
still far from equilibrium conditions. N2 fraction slightly increases as a [2] M.V. Twigg, London, Catalyst Handbook, 2nd ed., Manson Publishing Ltd., 1996.
[3] J. Kopyscinski, T.J. Schildhauer, S.M.A. Biollaz, Production of synthetic natural gas
result of the stoichiometry of the reaction. The experimental values (SNG) from coal and dry biomass—a technology review from 1950 to 2009, Fuel 89
measured at the reactor outlet agree well with the predicted values in (2010) 1763–1783.
in [4] M. Götz, J. Lefebvre, F. Mörs, A. McDaniel Koch, F. Graf, S. Bajohr, R. Reimert,
the minimum and maximum Wcat FCO 2
conditions considered, being the
T. Kolb, Renewable power-to-gas: a technological and economic review, Renew.
most noticeable deviation observed for hydrogen. It should be however
Energy 85 (2016) 1371–1390.
recalled that hydrogen fraction was determined through the molar [5] C.V. Miguel, M.A. Soria, A. Mendes, L.M. Madeira, Direct CO2 hydrogenation to
balance and it is therefore subject to the errors of the determination of methane or methanol from post-combustion exhaust streams—a thermodynamic
each species composition. study, J. Nat. Gas Sci. Eng. 22 (2015) 1–8.
[6] P. Lunde, F. Kester, Carbon dioxide methanation on a ruthenium catalyst, Ind. Eng.
Regarding to the reaction rate, it decreases steeply during the first Chem. Process Des. Dev. 13 (1974) 27–33.
half of the reactor length, particularly at low space velocities (i.e. high [7] E.Z. Golosman, V.N. Efremov, Industrial catalysts for the hydrogenation of carbon
in oxides, Catal. Ind. 4 (2012) 267–283.
Wcat FCO 2
), and then starts declining slowly until a value of 100
[8] T.T.M. Nguyen, L. Wissing, M.S. Skjøth-Rasmussen, High temperature methanation:
mmol⋅g cat⋅h-1 and 12.75 mmol⋅g cat⋅h-1 is reached at the exit for the catalyst considerations, Catal. Today 215 (2013) 233–238.
in
minimum and maximum Wcat FCO 2
considered, respectively. Both values [9] Y. Li, Q. Zhang, R. Chai, G. Zhao, Y. Liu, Y. Lu, Structured Ni-CeO2-Al2O3/Ni-foam
are very near to the observed reaction rates (Fig. 7b). catalyst with enhanced heat transfer for substitute natural gas production by syngas
methanation, ChemCatChem 7 (2015) 1427–1431.
The simulated methane yield profiles along the reactor length ob- [10] Z. Wan, J. Jiang, H. Ma, Y. Li, Y. Lu, F. Cao, A study of the heat transfer char-
tained in the temperature and reaction rate limits are shown in Fig. 8. acteristics of novel Ni-foam structured catalysts, Can. J. Chem. Eng. 94 (2016)
The model was able to capture the behaviour of the reactor perfor- 2225–2234.
[11] G.H. Watson, Methanation Catalysts, International Energy Agency Coal Research,
mance in the considered temperature window limits (250 °C–350 °C) London, 1980.
and in both of minimum and maximum reaction rates conditions. [12] Johnson Matthey, Chemical Catalysts: Catalysts for Olefin Processes, (2012).
[13] Clariant International Ltd, Catalysts for Syngas, (2010).
[14] Haldor Topsoe, PK-7R Low Temperature Methanation Catalyst, (2001).
5. Conclusions [15] Alvigo, Alvigo-Matros Catalysts, (2004).
[16] C.V. Miguel, M.A. Soria, A. Mendes, L.M. Madeira, A sorptive reactor for CO2
capture and conversion to renewable methane, Chem. Eng. J. 322 (2017) 590–602.
The kinetics of the methanation reaction over an industrial nickel- [17] J.-A. Dalmon, G.A. Martin, Intermediates in CO and CO2 hydrogenation over Ni
based catalyst was determined for the relevant 250 °C–350 °C tem- catalysts, J. Chem. Soc. Faraday Trans. 1: Phys. Chem. Condens. Phases 75 (1979)
perature window. Three reaction mechanisms assuming different in- 1011–1015.
[18] G.D. Weatherbee, C.H. Bartholomew, Hydrogenation of CO2 on group VIII metals:
termediate species were selected from the literature. The mechanism

135
C.V. Miguel et al. Journal of CO₂ Utilization 25 (2018) 128–136

II. Kinetics and mechanism of CO2 hydrogenation on nickel, J. Catal. 77 (1982) [26] G.F. Froment, K.B. Bischoff, Chemical Reactor Analysis and Design, 2nd ed., John
460–472. Wiley, New York, 1990.
[19] Si. Fujita, M. Nakamura, T. Doi, N. Takezawa, Mechanisms of methanation of [27] F. Ocampo, B. Louis, L. Kiwi-Minsker, A.-C. Roger, Effect of Ce/Zr composition and
carbon dioxide and carbon monoxide over nickel/alumina catalysts, Appl. Catal. A: noble metal promotion on nickel based CexZr1−xO2 catalysts for carbon dioxide
Gen. 104 (1993) 87–100. methanation, Appl. Catal. A: Gen. 392 (2011) 36–44.
[20] S.I. Fujita, N. Takezawa, Difference in the selectivity of CO and CO2 methanation [28] B.W. Wojciechowski, N.M. Rice, Experimental Methods in Kinetic Studies, Revised
reactions, Chem. Eng. J. 68 (1997) 63–68. edition, Elsevier Science B.V., Amsterdam, 2003.
[21] Q. Pan, J. Peng, T. Sun, S. Wang, S. Wang, Insight into the reaction route of CO2 [29] F.H.M. Dekker, A. Bliek, F. Kapteijn, J.A. Moulijn, Analysis of mass and heat
methanation: promotion effect of medium basic sites, Catal. Commun. 45 (2014) transfer in transient experiments over heterogeneous catalysts, Chem. Eng. Sci. 50
74–78. (1995) 3573–3580.
[22] Z.A. Ibraeva, N.V. Nekrasov, B.S. Gudkov, V.I. Yakerson, Z.T. Beisembaeva, [30] R.J. Berger, J. Pérez-Ramı́rez, F. Kapteijn, J.A. Moulijn, Catalyst performance
E.Z. Golosman, S.L. Kiperman, Kinetics of methanation of carbon dioxide on a testing: The influence of catalyst bed dilution on the conversion observed, Chem.
nickel catalyst, Theor. Exp. Chem. 26 (1991) 584–588. Eng. J. 90 (2002) 173–183.
[23] F. Koschany, D. Schlereth, O. Hinrichsen, On the kinetics of the methanation of [31] T. Van Herwijnen, H. Van Doesburg, W.A. De Jong, Kinetics of the methanation of
carbon dioxide on coprecipitated NiAl(O)x, Appl. Catal. B: Environ. 181 (2016) CO and CO2 on a nickel catalyst, J. Catal. 28 (1973) 391–402.
504–516. [32] M. Swickrath, M. Anderson, The development of models for carbon dioxide re-
[24] I. Kiendl, M. Klemm, A. Clemens, A. Herrman, Dilute gas methanation of synthesis duction technologies for spacecraft air revitalization, 42nd International
gas from biomass gasification, Fuel 123 (2014) 211–217. Conference on Environmental Systems, American Institute of Aeronautics and
[25] J. Horák, J. Pasek, London, Design of Industrial Chemical Reactors from Laboratory Astronautics, 2012.
Data, Heyden & Son Ltd., 1978. [33] S. Ergun, Fluid flow through packed columns, Chem. Eng. Prog. 48 (1952) 89–94.

136

You might also like