Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Journal of Environmental Chemical Engineering 9 (2021) 104629

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

CO2 conversion to methane using Ni/SiO2 catalysts promoted by Fe, Co


and Zn
Yan Resing Dias , Oscar W. Perez-Lopez *
Laboratory of Catalytic Processes–PROCAT, Department of Chemical Engineering, Federal University of Rio Grande do Sul (UFRGS), Ramiro Barcelos Street, 2777, CEP
90035-007, Porto Alegre, Brazil

A R T I C L E I N F O A B S T R A C T

Editor: GL Dotto Mitigation of CO2 emissions has become an important issue nowadays, as well as its reutilization as a carbon
source to obtain chemicals and fuels. CO2 conversion to CH4 comes up as an interesting alternative, since it is a
Keywords: natural gas substitute and H2 storage source. In this study, Ni/SiO2 catalysts promoted by Fe, Co and Zn were
Carbon dioxide conversion prepared by simple wet impregnation and evaluated in the CO2 conversion to methane. The samples were
Methanation
characterized by TGA-DTA, XRD, N2 physisorption, H2-TPR and TPO-DTA. The reactions were carried out at 1
Energy storage
atm in a fixed bed reactor with H2/CO2 = 4 and online GC analysis. The addition of a second metal enhanced the
Ni/SiO2 catalyst
Metal promotion surface area and lowered the Ni◦ crystallite size, indicating that metallic dispersion was improved. These
characteristics had positive influence on the methanation performance of the Fe and Co promoted catalysts,
which showed improved CO2 conversion and CH4 selectivity in comparison to unpromoted catalysts. Ni-Co/SiO2
presented the best results, attaining 73 % of CO2 conversion and 98.5 % of CH4 selectivity at 350 ◦ C and high
resistance to sintering. The catalyst promoted by Fe presented the higher resistance to carbon formation, whereas
the promotion with Zn led to a strong decrease in selectivity for CH4 and a consequent increase in selectivity for
CO. These results suggest the use of both Fe and Co as promoters of Ni/SiO2 catalysts has a high potential for the
CO2 conversion to CH4.

1. Introduction hydrogen-rich fuels as CH4, with high H/C ratio, becomes an important
route to H2 storage, since the H2 obtained from water electrolysis using
Carbon dioxide (CO2) emissions have been a major threat to Earth’s electricity generated from renewable sources – known as power-to-gas
environment since its emissions increase year by year [1,2]. The processes – can be difficult to store and transport, whereas fuels as CH4
greenhouse effect, which causes the global warming, increases the can be easily and safely transported in existing natural gas pipelines [6,
temperature of the planet, melts glaciers and rises sea levels, etc., is 13].
mainly attributed to these CO2 emissions, which correspond to 80 % of One of the recently most studied routes to convert CO2 via catalytic
total emissions [3–5]. The origin of these large CO2 emissions comes hydrogenation processes is the methanation or Sabatier reaction [2,
from the extensive use of carbon-based compounds, since oil and coal 14–18]. This reaction is highly exothermic, thus occurring at relatively
are still the major fuels and energy sources. In addition, these are finite low temperatures, between 200 and 450 ◦ C, and can be carried out at
resources, thus forcing the search and use of alternative and renewable atmospheric pressure, although high pressures leads to higher CO2
sources, such as wind, biomass, and solar energy [6,7]. conversion [19]. The proper H2/CO2 ratio must be selected, as ratios
The carbon capture and utilization (CCU) is gaining increasing equal or higher than the stoichiometric favors CO2 consumption towards
attention because it allows the reuse of CO2 as a raw material to the CH4 formation [20]. The kinetic barriers imposed by thermodynamics
synthesis of fuels and chemicals, such as methane, syngas, methanol and and difficult CO2 activation due to its high stability are challenges in CO2
higher alcohols, formic acid, hydrocarbons, etc. [5,8–10]. Among these methanation [21].
products, methane (CH4) stands out as a feedstock to chemical synthesis, Two mechanisms of CO2 methanation are usually described in the
mainly as synthetic natural gas (SNG), acting as a substitute or com­ literature. Firstly, the direct methanation, which occurs according to
plement to natural gas sources [7,11,12]. In addition, CO2 conversion to Equation (1) with nominal ratio H2:CO2 = 4:1 and can be associated

* Corresponding author.
E-mail address: perez@ufrgs.br (O.W. Perez-Lopez).

https://doi.org/10.1016/j.jece.2020.104629
Received 14 September 2020; Received in revised form 5 October 2020; Accepted 13 October 2020
Available online 19 October 2020
2213-3437/© 2020 Elsevier Ltd. All rights reserved.
Y.R. Dias and O.W. Perez-Lopez Journal of Environmental Chemical Engineering 9 (2021) 104629

with the formation of intermediaries as formates and carbonates [22, promoter [43–45]. Transition metals, such as Fe, Co, and Cu, lantha­
23]. The second and most accepted is the path of CO as intermediary, nides as La and Ce, alkaline and alkaline earth metals as Na, K, Mg, and
where CO2 is firstly transformed to CO through Equation (2) and then Ca, have been commonly chosen as promoters [46–48]. Liu et al. pre­
hydrogenated to methane according to Equation (3) [11,19]. Equation pared 10 %Ni-3 %Co catalysts supported over alumina and ordered
(2) is the reverse water-gas shift (RWGS) reaction. mesoporous alumina (OMA) by a one-pot method. The best results were
obtained for Ni-Co/OMA in comparison to the monometallic and/or
4H2 + CO2 ⇆ CH4 + 2H2O ΔH0R = − 165 kJ mol− 1
(1) alumina-supported catalysts, with 78 % CO2 conversion and 99 %
H2 + CO2 ⇆ CO + H2O ΔH0R = 41 kJ mol− 1
(2) selectivity for methane at 400 ◦ C, which was assigned to the synergistic
effect between Ni and Co. The use of OMA as support has increased the

3H2 + CO ⇆ CH4 + H2O ΔHR = − 206 kJ mol − 1
(3) surface area and resulted in small crystallite sizes, whereas the presence
of Co has enhanced H2 chemisorption, even though these catalysts
Catalysts for CO2 methanation need to have high activity and high
presented lower basicity [49]. A similar approach was used by Xu et al.
selectivity for CH4 at low temperatures, as well as resistance to deacti­
with several Ni-Co/OMA catalysts and total metal loading of 10 %. The
vation. Noble metals such as Rh, Ru, and Pd are very active and selective
authors found that a 20 % Co/(Co + Ni) catalyst had the best activity
to methanation. However, they are expensive, which limits their utili­
and selectivity, when compared to monometallic or high Co loading
zation [24–28]. Therefore, transition metals have been widely applied as
catalysts, because of the strong Ni-Co synergistic effect [50]. Mog­
the active phase, such as Ni, Fe and Co. In spite of its susceptibility to
haddam et al. obtained Ni/Al2O3 by the one-pot sol-gel method pro­
sintering and carbon deposition, Ni catalysts stand out as the most
moted by Fe, Co, Cu, Zr and La, with 30 % of Ni and 5 % of promoter. It
commonly chosen ones due to their high activity and selectivity at low
was observed that the Fe-promoted catalyst showed the best perfor­
temperatures [5,29,30]. Although sintering is favored at high temper­
mance with approximately 71 % of CO2 conversion and 99 % selectivity
atures, that are not usually seem in CO2 methanation, it may occur at
for methane at 350 ◦ C, which could be attributed to the Ni-Fe alloy
low temperatures due to the exothermic nature of the reaction. There­
improving H2 adsorption and CO2 dissociation. Moreover, enhancing Fe
fore, characteristics such as high metal-support interaction, small Ni
content from 5 to 7 % increased the activity at lower temperatures,
crystallites and high dispersion can improve resistance to sintering [31,
accompanied by a higher reducibility and great stability in a 10 h test
32]. These properties are also desirable to avoid carbon deposition,
[43]. Multiple promoted Ni/Al2O3 catalysts were synthesized by Liang
along with surface basicity that can increase chemisorption and acti­
et al. by incipient impregnation, from alkali to transition metals When
vation of CO2 making the active sites unavailable to coking [31–33].
compared to unpromoted catalysts, the Zn based catalysts showed better
Most of the methanation catalysts are prepared by impregnation over
results at low temperatures – although not so expressive as Fe and Co
supports, such as Al2O3, SiO2, CeO2, ZrO2, and TiO2, that provide
based catalysts – due to a weakening effect on the metal-support inter­
important characteristics, e.g., surface area and porosity, thermal and
action [47].
chemical stability, among others [27,34–36]. The effect of solution pH
A few works have focused on Ni/SiO2 promoted catalysts for the CO2
on the adsorption of the metal precursor depends on the point of zero
methanation. Pandey et al. tested Ni-Fe catalysts supported over SiO2
charge (PZC) of the support as it may create surface hydroxyl groups that
(total metal loading of 10 %) and Al2O3 (total metal loading of 10 % and
acts as adsorption sites, which affect the particle size of metals and the
30 %), with variable Ni/Fe ratios. The optimal Ni content was 75 % of
interaction between metal and support [37,38]. The drying step is also
total metal loading over alumina and silica, in which alumina was the
fundamental, as the solvent removing may carry the precursor from the
best support, although those supported over silica were more easily
pores to the external surface of the support, which can result in
reduced. The Ni-Fe alloy presented a synergistic effect resulting in a
agglomeration of precursor in larger crystallites and even block the
considerable amount of metal sites that boosted activity and CH4 yield.
pores of the support [38,39].
The same authors also studied the effect of different supports, namely
Nickel catalysts were significantly studied in CO2 methanation,
Al2O3, SiO2, ZrO2, TiO2 and Nb2O5, on Ni-Fe systems, with 10 % of total
especially over alumina. In order to evaluate the influence of the support
metal loading. As previously, 75 % of Ni in the total metal loading
in catalysts properties and performance, Le et al. impregnated 10 % Ni
supported over alumina had the best results, with a similar behavior to
over different supports, namely γ-Al2O3, SiO2, CeO2, ZrO2, and TiO2 and
the other supports, except for Nb2O5, which formed niobate compounds
found that the Ni/CeO2 had the best results, attaining 100 % of CO2
with low activity. Excluding Nb2O5, silica showed the lowest improve­
conversion and CH4 yield at temperatures below 250 ◦ C, which was
ments on activity and selectivity, which could be related to a lower CO2
ascribed to smaller crystallite sizes [40]. All catalysts eventually reached
adsorption capability of this support [51,52]. Ni-Co/SiO2 catalysts
100 % CO2 conversion at higher temperatures. Zhang et al. tested 5 %
prepared by wet impregnation with a total metal loading of 10 % were
and 20 % Ni on Al2O3 prepared by impregnation, with the support
used by Guo and Lu. The authors found that with low Co additions,
previously calcined at 600, 800 or 1000 ◦ C, observing that higher
under 4 %, the catalyst activity was promoted, as well as the reducibility
calcination temperatures lower the interaction of Ni-Al and provide
[53]. Wu et al. prepared Fe-Ni catalysts by hydrothermal method sup­
more active sites for the reaction. Higher Ni loadings show better ac­
ported in a hydrophobic SiO2 for syngas methanation. The authors
tivity at low temperatures even with poorer metal dispersion [41].
observed that the CO conversion was higher for Fe/Ni ratios equal to or
Co-precipitated catalysts were also designed, such as the Ni-Al2O3
higher than 1, although CH4 selectivity presented opposed results, also
ultrasound-assisted catalysts prepared by Daroughegi et al., with Ni
showing that C2-C4 compounds were significantly formed at low tem­
loadings between 15 and 33 %. It was shown that up to 25 % of Ni, both
peratures. The Fe1.5Ni15 catalyst attaining 95.2 % of CO conversion
CO2 conversion and CH4 selectivity are enhanced to 80 % and 100 %,
and 68.7 % of CH4 selectivity at 320 ◦ C [54].
respectively, due to the higher surface area and improved reducibility
Although Al2O3 is knowingly a good support to CO2 methanation
that are related to the creation of extra active sites, also reaching high
catalysts, it has a high interaction with the impregnated metals, which
stability over a 10 h test. The influence of GHSV and H2/CO2 ratio were
frequently difficult the catalysts reducibility mainly due to the formation
also evaluated. It was observed that increasing GHSV decreased the
of spinel phases that are inactive in the reaction [43]. SiO2 supported
catalytic activity, whereas higher H2/CO2 ratios improved the catalytic
catalyst were barely studied for CO2 methanation. This support pre­
performance [42].
senting lower interaction with metals than Al2O3, avoiding formation of
In order to modify characteristics such as metal dispersion, enhanced
mixed inactive compounds and reflecting positively in reducibility [43,
basic sites and CO2 adsorption capability, reducibility, resistance to
55]. In a previous work, we studied the influence of Cu promotion on
deactivation, etc., and consequently activity and selectivity, some works
activity, selectivity, and resistance to deactivation on Ni/SiO2 catalysts
have focused on bimetallic catalysts, in which the second metal acts as a
in the CO2 conversion to CH4 [56]. Copper has promoted Ni dispersion,

2
Y.R. Dias and O.W. Perez-Lopez Journal of Environmental Chemical Engineering 9 (2021) 104629

surface area and reducibility, and lowered the crystallite size of the CO2 conversion and selectivity for CH4 and CO (on a dry basis) were
catalysts. Although it presented high stability and resistance to sintering calculated according to Eq.s 5–7 [56]:
and carbon deposition, the activity and CH4 selectivity were lower than
FCO2 in − FCO2 out
unpromoted catalysts [56]. In this context, this work proposes the XCO2 (%) = × 100 (5)
FCO2 in
evaluation of the transition metals Fe, Co, and Zn as promoters of
Ni/SiO2 catalyst, aiming to evaluate its influence in the surface prop­ FCH4 out
erties and improve the activity, selectivity and resistance to deactivation SCH4 (%) = × 100 (6)
FCH4 out + FCOout
in the CO2 methanation.
FCOout
2. Experimental SCO (%) = × 100 (7)
FCH4 out + FCOout

2.1. Catalyst preparation 3. Results and discussion

The catalysts were prepared by wet impregnation according to pre­ 3.1. Catalyst characterization
viously published works [56,57]. Nickel nitrate (Ni(NO3)2.6H2O), and
the respective nitrate of the promoter (Fe(NO3)3.9H2O, Co(NO3)2.6H2O Thermogravimetric analysis shows that decomposition of the un­
or Zn(NO3)2.6H2O) were dissolved in 50 mL of deionized water, and calcined samples occurs in several steps, according to the three thermal
then 1 g of SiO2 (Aerosil® 200, Evonik Degussa) was added. The mixture events shown in Fig. 1(a) and confirmed by the DTG exothermic peaks
was stirred for 4 h at ambient temperature. The samples were dried in an presented in Fig. 1(b). These events are in agreement with the results
oven, and then calcined under air flow at 500 ◦ C for 2 h. The Ni content reported in the literature [67,68].
was 10 % and the promoter content equal to 1 %, based on our previous Table 1 contains the data obtained from the N2 physisorption. The
work [56,57]. The samples were identified as Ni-M/SiO2, where M is the addition of promoters shows a significant enhancement in specific sur­
metal promoter Fe, Co or Zn. face area and pore volume when compared to Ni10, especially for
Ni10Fe1 and Ni10Zn1, which also showed lower pore diameters.
2.2. Characterization of the catalysts Although Ni10Co1 had the highest pore diameter, the increasing in pore
volume led to a specific surface area higher than that of the Ni10. All
The XRD patterns of the samples were obtained in a D2 Phaser samples showed type IV(a) isotherms, which is characteristic of meso­
(Bruker) diffractometer, with a Cu-Kα radiation source at 30 kV and 10 porous materials with pore condensation, and confirmed by the pore
mA [56,57]. The crystallite size was obtained using the Scherrer’s [58, diameter distribution (Figure S1 in supplementary material). The iso­
59]. therms of these catalysts also present a type H2(b) hysteresis, that may
The specific surface area of the samples was measured by N2 phys­ indicate a pore blockage during condensation-evaporation process in
isorption in a Quantachrome 4200e analyzer. The samples were high relative pressure in pores of large neck width [69,70].
degassed at 300 ◦ C for 3 h under vacuum. The adsorption and desorption Fig. 2 shows the XRD patterns of the calcined and reduced samples.
tests were carried out using N2 at − 196 ◦ C. The specific surface area was The calcined samples exhibited a broad peak near 22◦ related to amor­
determined using the multipoint BET method, whereas pore size and phous silica. The peaks at 37.3, 43.5, and 63.1◦ indicate the presence of
pore volume were calculated via the BJH method [60,61]. NiO phase (Fig. 2a). For the reduced catalysts, Fig. 2b shows reflections
TGA-DTG and temperature programmed oxidation (TPO) analysis at 44.6 and 52◦ assigned to the Ni0 phase [56,67].
were carried out in a thermobalance SDT Q600, TA Instruments, under Peaks relative to metal promoters or mixed nickel-promoters phases
an air flow rate of 100 mL.min− 1 and a heating rate of 10 ◦ C.min− 1. TGA- were not observed, possibly due to the low promoter loadings used [71].
DTG was performed for the uncalcined samples from room temperature The Ni10 catalyst had the most intense peaks, and consequently largest
to 500 ◦ C, whereas TPO was carried out using the spent samples, from crystallite average size (Table 1) for both calcined and reduced samples,
room temperature to 800 ◦ C [62,63]. The amount of carbon deposited whereas the peak intensity of the promoted catalysts was lower, which
was quantified from TPO profiles according to Eq. 4: indicates that the promotion led to less crystallinity and lower crystallite
( ) sizes, except for the reduced Ni10Zn1 sample, as presented in Table 1.
mzone III f − mzone III i
C mg/ = x 10 (4)
gcat mcat
The reducibility of the calcined samples was evaluated by H2-tem­
perature programmed reduction (H2-TPR). The samples were degassed
at 100 ◦ C under N2 flow, and then heated at 10 ◦ C.min− 1 up to 800 ◦ C,
with a mixture of 5 % H2/N2 (30 mL.min− 1) [64,65].

2.3. Catalytic activity

A detailed description of the procedure can be found in [48]. Briefly,


the experiments were carried out at 1 atm in a tubular quartz reactor. In
each test, approximately 100 mg of catalyst were reduced in situ at 400

C for 1 h in 10 % H2/N2 at a flow rate of 100 mL.min− 1. The catalytic
tests were carried out with a mixture of CO2/H2/N2 with a molar ratio
1/4/15 and total flow rate of 100 mL.min− 1. The gas flow rates were
adjusted by digital mass flow controllers. Experiments where performed
in a stepwise mode within the temperature range of 200–400 ◦ C. Ex­
periments with a fixed temperature of 400 ◦ C and time-on-stream of 5 h
were also carried out. The analysis of the products was accomplished by
on-line gas chromatography (Varian 3600Cx, Porapak Q column) with a Fig. 1. Thermogravimetric analysis of the uncalcined Ni-M/SiO2 (M = Fe, Co,
thermal conductivity detector and N2 as the carrier gas [56,66]. Zn) samples: TG (a) and DTG (b).

3
Y.R. Dias and O.W. Perez-Lopez Journal of Environmental Chemical Engineering 9 (2021) 104629

Table 1 Table 2 shows that the area of the first peak decreases in the
Surface properties and crystallite size of the Ni-M/SiO2 (M = Fe, Co, Zn) following order: Ni10 > Ni10Fe1 > Ni10Co1 > Ni10Zn1. Therefore, the
catalysts. area of the peaks at higher temperatures increases, indicating an in­
Surface properties Crystallite size (nm) crease in the amount of small nickel particles, and consequently smaller
SBET Vpore Dpore Calcined Reduced After
average crystallite sizes, which agrees with the results of XRD. Among
Sample (m2 (cm3 (nm)b at 500 ◦ C at 400 ◦ C reaction the promoted samples, those with Fe or Co showed higher reduction
g− 1)a g− 1)b (DNiO)c (DNi◦ )d (D0Ni temperatures of the second and third peaks, which indicates that these
d
spent) metals enhance Ni dispersion by presenting smaller crystallites that are
Ni10 153.0 0.352 16.4 14.4 15.3 16.0 more difficultly reducible and consequently requiring higher tempera­
Ni10Fe1 176.3 0.458 12.9 10.4 12.5 13.5 tures. On the other hand, the sample promoted by Zn had the lowest
Ni10Co1 165.3 0.405 18.1 11.2 11.4 11.8 reduction temperatures of the second and third peaks and, at the same
Ni10Zn1 174.5 0.456 12.8 11.8 17.1 16.1
time, the largest area of the second peak among all samples.
a
By the BET method.
b
By the BJH method.
c
From XRD reflection of NiO at 43.5◦ . 3.2. Catalytic tests
d
From XRD reflection of Ni0 at 44.6◦ .
Fig. 4 displays the CO2 conversion values versus reaction tempera­
This could also indicate the increasing on metal dispersion that is in ture, as well as the equilibrium conversion based on results reported by
accordance to the surface area results. However, among the reduced Schaaf et al. [77]. At 200 ◦ C, almost no activity was observed, except for
samples, Ni10Zn1 showed an expressive increase of crystallite size, the Ni10Zn1, due to the low reaction temperature and therefore low
which may indicate a sintering occurrence during the activation step. reaction rate. The CO2 conversion increases significantly with the tem­
Fig. 3 shows the H2-TPR curves obtained for calcined samples. The perature increase until 350 ◦ C, reaching the conversions of 53 %, 69 %,
Ni10 catalyst shows two main reduction peaks. The first peak is related and 73 % for the Ni10, Ni10Fe1 and Ni10Co1 catalysts, respectively.
to the reduction of large bulk NiO (Ni2+) crystallites to Ni◦ with low These results show that Fe and Co as promoters lead to higher activity,
interactions with the support, whereas the second peak is ascribed to the possibly due to a synergetic effect between Ni and these metals [49]. In
reduction of small NiO crystallites to Ni◦ with stronger interactions with addition, the larger surface area and pore volume, enhanced reducibility
the support [25,56,72]. of bulk NiO, and lower crystallite sizes may have contributed to
The Fe, Co, and Zn-promoted catalysts show three reduction peaks, increasing the activity of Ni10Fe1 and Ni10Co1 in comparison to Ni10.
according to the deconvolution data (Table 2). The addition of the Although the Zn-promoted catalyst showed high surface area and pore
second metal modified the reducibility of NiO. The peaks related to the volume along with enhanced reducibility of bulk NiO, as Ni10Fe1 and
reduction of bulk NiO were shifted to lower temperatures, whereas the Ni10Co1, it had a different behavior regarding activity, with higher
peaks related to small NiO crystallites were shifted to higher tempera­ initial activity at 200 ◦ C but lower activity at temperatures over 250 ◦ C,
tures. For the Zn-promoted catalyst, it was also observed a splitting of attaining 52 % of conversion at 400 ◦ C. The high activity at 200 ◦ C can
the second reduction peak into two peaks at 465 and 541 ◦ C, which be ascribed to the larger crystallite size (Table 1) due to its high
indicates that Zn possibly had a strong interaction with Ni and decreased reducibility, while the behavior at temperatures over 250 ◦ C could be
the metal-support interaction, possibly due to the formation of Ni-Zn related to an inactive Ni-Zn alloy to hydrogenation or to Zn activity in
alloys, enhancing the reducibility of smaller particles, although ZnO is the WGS reaction that occurs at the same temperature range and pro­
not reducible [47]. Fe and Co additions increased the reducibility of duces CO2 [78]. On the other hand, Ni10, Ni10Fe1 and Ni10Co1 show a
larger NiO crystallites, whereas an opposite effect was observed in CO2 conversion decrease at 400 ◦ C that occurs due to a thermodynamic
smaller NiO crystallites, in which the reduction peak was slightly shifted limitation because the RWGS reaction is favored at this condition. This
to higher temperatures and split into two peaks. The second metal decrease can also be due to catalyst sintering or carbon deposition over
addition improves the reducibility of NiO possibly due to a Ni-M syn­ the catalysts [73].
ergetic and spillover effect, or even attenuation in metal-support inter­ Fig. 5 shows the CH4 and CO selectivity with its respective average
action, as suggested by previous works [47,73]. standard deviation. At 200 ◦ C, these selectivities were of 100 and 0 %,
Bulk Fe and Co reduction peaks were not observed because of the low respectively, for all catalysts. These results indicate that only the direct
promoter amount and possibly due to overlapping by the NiO reduction methanation mechanism occurs at this temperature, whereas CO starts
or even the presence of Ni-Fe and Ni-Co alloys. The second and third to be formed over 250 ◦ C. The promoted samples Ni10Fe1 and Ni10Co1
reduction peaks for both catalysts may also be related to the reduction of attained CH4 selectivity over 90 %, whereas the unpromoted Ni10 pre­
Fe and Co oxides [54,74–76]. sents CH4 selectivity near to 90 % only at temperatures above 300 ◦ C. In
the same way to the CO2 conversion, the CH4 selectivity decreased at

Fig. 2. X-ray diffraction patterns of Ni-M/SiO2 (M = Fe, Co, Zn) samples: (a) calcined samples and (b) reduced samples, where (◆) SiO2, (*) NiO and (◦ ) Ni0.

4
Y.R. Dias and O.W. Perez-Lopez Journal of Environmental Chemical Engineering 9 (2021) 104629

400 ◦ C, whereas the CO selectivity increased, which indicates that more


CO was been produced due to the RWGS reaction.
The Zn-promoted catalyst presented a different trend after 200 ◦ C,
showing higher selectivity to CO until 350 ◦ C and reaching a maximum
of 67 % at 250 ◦ C, whereas the CH4 selectivity increases at 400 ◦ C. This
indicates that Zn promote the CO formation, possibly due to its activity
in the WGS reaction and therefore in the RWGS reaction [78].
In Figure S2, the CH4 selectivity as a function of CO2 conversion was
plotted. The yield for methane and the yield for CO are shown in
Table S1 and Table S2, respectively. Figure S2 shows that the catalysts
Ni10, Ni10Fe1 and Ni10Co1 presented a higher CH4 selectivity than
Ni10Zn1 for the same CO2 conversion range, from 10 to 55 %. In
addition, the superiority of the Ni10Fe1 and Ni10Co1 catalysts over
Ni10 and Ni10Zn1 is noted for high CO2 conversions. This difference is
more evident when comparing the CH4 yield in Table S1, where the
highest yields were obtained at temperatures of 350 and 400 ◦ C, for the
Ni10Fe1 and Ni10Co1 catalysts. At 350 ◦ C the methane yield was in the
following order: Ni10Co1 > Ni10Fe1 > Ni10 > > Ni10Zn1. On the other
hand, Table S2 shows that the highest yields for CO were obtained with
the Ni10Zn1 catalyst, reaching a CO yield of 19 % at 350 ◦ C.
Fig. 6 shows the catalytic tests performed at 400 ◦ C. All catalysts
exhibited high stability along a time-on-stream of 5 h, with Ni10Zn1 at a
low CO2 conversion level, which agrees with the results obtained in the
stepwise mode within the temperature range of 200–400 ◦ C. These re­
sults demonstrate that the unpromoted catalyst has high resistance to
deactivation as the promoted catalysts. Also, the Fe promotion seems to
maintain the CO2 conversion nearly constant, indicating a stabilizing
effect on Ni◦ crystallites [71].

3.3. Catalyst characterization after the reactions

Fig. 3. Deconvolution of H2-TPR profiles of the calcined Ni-M/SiO2 (M = Fe, The XRD patterns of the catalysts after the reactions are presented in
Co, Zn) samples. Fig. 7. They show the same phases present in the reduced catalysts
before the reaction, with a broad peak at 22◦ related to amorphous SiO2
and peaks at 44.6◦ and 52◦ related to Ni0. Carbon deposits were not
Table 2
identified in the patterns due to the overlap with the SiO2 peak and to
Temperatures and peak areas of the Ni-M/SiO2 (M = Fe, Co, Zn) catalysts after
the low quantity formed [57]. The crystallite sizes (Table 1) had no
the TPR profile deconvolution.
significant change before and after the reaction, which indicates a high
Temperature (◦ C) Relative area (%) resistance to sintering, where the Co-promoted catalyst remained with
Total area
Sample 1st 2nd 3rd 1st 2nd 3rd the smallest crystallite size, around 11 nm.
(a.u.)
peak peak peak peak peak peak Fig. 8 shows the TPO traces of the spent catalysts, which can be
Ni10 390 509 – 83.98 16.02 – 874.6 separated in three zones of weight changes. The first one (i), between
Ni10Fe1 386 513 556 82.30 13.53 4.17 867.0 ambient temperature and 200 ◦ C, corresponds to a water loss due to
Ni10Co1 370 529 582 70.03 20.37 9.60 718.6
moisture evaporation, and the second zone (ii) shows the oxidation of
Ni10Zn1 373 465 541 55.30 35.69 9.01 747.4
the metallic particles from 200 to 450 ◦ C. According to the weight
variation in this region, Ni10Fe1 showed the largest mass increase,
which could indicate that this catalyst has the largest amount of metallic
Ni after the reaction. The weight losses in Zone (iii), within the tem­
perature range of 450− 800 ◦ C, could be assigned to carbon oxidation,
which increases in the following sequence: Ni10Fe1 < Ni10 ~ Ni10Zn1
< Ni10Co1. However, these losses were very low (Table 3). These results
show that Fe promotion has enhanced the resistance to carbon deposi­
tion, since the Ni10Fe1 catalyst had the lowest carbon formation,
regarding the low carbon presented on each catalyst.
Table 4 shows some results reported in the literature with the cata­
lysts evaluated in this work. Liu et al. and Xu et al. obtained better re­
sults regarding the CO2 conversion in comparison to Ni10 when
supported by ordered mesoporous alumina (OMA), but the CH4 selec­
tivity was very close. Ni10Co1 attained similar results as those of
Moghaddam et al., despite the lower amount of metals used in the
present work. The Ni10Fe1 catalyst had results at the same level as those
of Moghaddam et al., and was even better than the catalysts of Wierz­
bicki et al. and Daroughegi et al. However, the catalyst promoted by Zn
had results worse than those of Liang et al. The differences among these
Fig. 4. CO2 conversion as a function of the reaction temperature for the Ni-M/ results can be assigned to those catalysts being supported over alumina,
SiO2 (M = Fe, Co, Zn) catalysts. and also due to the higher GHSV used in this work [43,47,49,50,73,79].

5
Y.R. Dias and O.W. Perez-Lopez Journal of Environmental Chemical Engineering 9 (2021) 104629

Fig. 5. Selectivity to CH4 (a) and CO (b) as a function of the reaction temperature for the Ni-M/SiO2 (M = Fe, Co, Zn) catalysts.

Fig. 8. Weight variation obtained by TPO of the spent Ni-M/SiO2 (M = Fe, Co,
Fig. 6. CO2 conversion with the time on stream at 400 ◦ C. Zn) catalysts.

Table 3
Weight variation and carbon formation obtained by TPO of the spent Ni-M/SiO2
(M = Fe, Co, Zn) catalysts.
Catalyst Zone (i) Zone (ii) Zone (iii) Carbon production (mg/
(%) (%) (%) gcat)

Ni10 − 2.3 +1.47 − 0.70 7.0


Ni10Fe1 − 2.1 +1.69 − 0.58 5.8
Ni10Co1 − 2.0 +1.30 − 1.17 11.7
Ni10Zn1 − 2.8 +1.46 − 0.73 7.3

selectivity to methane, influenced by a higher surface area, lower crys­


tallite size and enhanced reducibility of Ni, whereas the promotion by Zn
decreased the CO2 methanation and increased the selectivity to CO due
to a WGS/RWGS reaction favoring. Based on the CH4 yield, the best
results were obtained at 350 ◦ C in the following order: Ni10Co1 >
Fig. 7. Diffractograms of the spent Ni-M/SiO2 (M = Fe, Co, Zn) catalysts, where
Ni10Fe1 > Ni10 > > Ni10Zn1.
(◆) SiO2 and (◦ ) Ni0.
Ni/SiO2 catalysts promoted by Co or Fe showed high potential for the
CO2 methanation. The addition of Co increases the catalyst sintering
4. Conclusions resistance, whereas Fe increases carbon deposition resistance, which
points to the use of both metals together as promoters.
The addition of a second metal (Fe, Co or Zn) increases the specific
surface area and pore volume of the Ni/SiO2 catalyst. The addition of Co CRediT authorship contribution statement
or Fe improved the dispersion of Ni, leading to a smaller crystallite size
and altering the reducibility of NiO, which forms particles with higher Yan Resing Dias: Methodology, Validation, Investigation, Writing -
interaction with SiO2. On the other hand, the addition of Zn increased original draft. Oscar W. Perez-Lopez: Conceptualization, Writing - re­
the reducibility of NiO, forming larger particles of Ni0. view & editing, Supervision.
The addition of Fe and Co enhanced the CO2 conversion and

6
Y.R. Dias and O.W. Perez-Lopez Journal of Environmental Chemical Engineering 9 (2021) 104629

Table 4 synthetic natural gas, RSC Adv. 6 (2016) 28489–28499, https://doi.org/10.1039/


Comparison of some works reported in the literature for CO2 methanation. c6ra01139j.
[6] B. Lee, H. Lee, S. Kang, H. Lim, Stochastic techno-economic analysis of power-to-
Catalyst Temperature H2/ GHSV XCO2 SCH4 References gas technology for synthetic natural gas production based on renewable H2 cost
(◦ C) CO2 (mL g−cat1 (%) (%) and CO2 tax credit, J. Energy Storage. 24 (2019) 100791, https://doi.org/10.1016/
in h− 1)a j.est.2019.100791.
feed [7] S. Walspurger, G.D. Elzinga, J.W. Dijkstra, M. Sarić, W.G. Haije, Sorption enhanced
methanation for substitute natural gas production: experimental results and
10 %Ni/ 350 4 60000 53.5 96 This work thermodynamic considerations, Chem. Eng. J. 242 (2014) 379–386, https://doi.
SiO2 org/10.1016/j.cej.2013.12.045.
10 %Ni-1 350 4 60000 69 97 This work [8] M.D. Esrafili, L. Dinparast, A DFT study on the catalytic hydrogenation of CO2 to
%Fe/ formic acid over Ti-doped graphene nanoflake, Chem. Phys. Lett. 682 (2017)
SiO2 49–54, https://doi.org/10.1016/j.cplett.2017.06.011.
10 %Ni-1 350 4 60000 73 98.5 This work [9] S. Xiong, Y. Lian, H. Xie, B. Liu, Hydrogenation of CO2 to methanol over Cu/ZnCr
%Co/ catalyst, Fuel 256 (2019), https://doi.org/10.1016/j.fuel.2019.115975, 115975.
[10] R.E. Owen, P. Plucinski, D. Mattia, L. Torrente-Murciano, V.P. Ting, M.D. Jones,
SiO2
Effect of support of Co-Na-Mo catalysts on the direct conversion of CO2 to
10 %Ni-1 400 4 60000 52 82 This work
hydrocarbons, J. CO2 Util. 16 (2016) 97–103, https://doi.org/10.1016/j.
%Zn/
jcou.2016.06.009.
SiO2 [11] E. Giglio, A. Lanzini, M. Santarelli, P. Leone, Synthetic natural gas via integrated
10 %Ni/ 400 4 10000 70 95 [49] high-temperature electrolysis and methanation: part I-Energy performance,
OMA J. Energy Storage. 1 (2015) 22–37, https://doi.org/10.1016/j.est.2015.04.002.
10 %Ni/ 350 4 15000 65 97 [50] [12] E.I. Koytsoumpa, S. Karellas, Equilibrium and kinetic aspects for catalytic
OMA methanation focusing on CO2 derived Substitute Natural Gas (SNG), Renew. Sust.
30 %Ni-5 350 3.5 9000 71 99 [43] Energy Rev. 94 (2018) 536–550, https://doi.org/10.1016/j.rser.2018.06.051.
%Fe/ [13] F.D. Meylan, F.P. Piguet, S. Erkman, Power-to-gas through CO2 methanation:
Al2O3 assessment of the carbon balance regarding EU directives, J. Energy Storage. 11
20 %Ni- 350 4 12000 62 98 [73] (2017) 16–24, https://doi.org/10.1016/j.est.2016.12.005.
2.8 % (h− 1) [14] H. Lu, X. Yang, G. Gao, J. Wang, C. Han, X. Liang, C. Li, Y. Li, W. Zhang, X. Chen,
Fe-HT Metal (Fe, Co, Ce or La) doped nickel catalyst supported on ZrO2 modified
mesoporous clays for CO and CO2 methanation, Fuel 183 (2016) 335–344, https://
25 %Ni-5 350 3.5 9000 58 98 [79]
doi.org/10.1016/j.fuel.2016.06.084.
%Fe-
[15] S. Tada, S. Ikeda, N. Shimoda, T. Honma, M. Takahashi, A. Nariyuki, S. Satokawa,
Al2O3 Sponge Ni catalyst with high activity in CO2 methanation, Int. J. Hydrogen Energy
10 %Ni-3 400 4 10000 78 99 [49] 42 (2017) 30126–30134, https://doi.org/10.1016/j.ijhydene.2017.10.138.
%Co/ [16] T.A. Le, T.W. Kim, S.H. Lee, E.D. Park, Effects of Na content in Na/Ni/SiO2 and Na/
OMA Ni/CeO2 catalysts for CO and CO2 methanation, Catal. Today 303 (2018) 159–167,
8 %Ni-2 350 4 15000 72 99 [50] https://doi.org/10.1016/j.cattod.2017.09.031.
%Co/ [17] A. Kim, D.P. Debecker, F. Devred, V. Dubois, C. Sanchez, C. Sassoye, CO2
OMA methanation on Ru/TiO2 catalysts: On the effect of mixing anatase and rutile
30 %Ni-5 350 3.5 9000 70 99 [43] TiO2supports, Appl. Catal. B Environ. 220 (2018) 615–625, https://doi.org/
%Co/ 10.1016/j.apcatb.2017.08.058.
Al2O3 [18] P. Sabatier, J.B. Senderens, New synthesis of methane, Comptes Rendus. 134
20 %Ni-5 350 4 16000 68 62b [47] (1902) 514–516.
[19] K. Stangeland, D. Kalai, H. Li, Z. Yu, CO2Methanation: The Effect of Catalysts and
%Zn/ (h− 1)
Reaction Conditions, Energy Procedia 105 (2017) 2022–2027, https://doi.org/
Al2O3
10.1016/j.egypro.2017.03.577.
a
GHSV denoted in these units unless indicated. [20] J. Gao, Y. Wang, Y. Ping, D. Hu, G. Xu, F. Gu, F. Su, A thermodynamic analysis of
b methanation reactions of carbon oxides for the production of synthetic natural gas,
Value expressed in terms of CH4 yield.
RSC Adv. 2 (2012) 2358–2368, https://doi.org/10.1039/c2ra00632d.
[21] A. Álvarez, M. Borges, J.J. Corral-Pérez, J.G. Olcina, L. Hu, D. Cornu, R. Huang,
Declaration of Competing Interest D. Stoian, A. Urakawa, CO2Activation over Catalytic Surfaces, ChemPhysChem. 18
(2017) 3135–3141, https://doi.org/10.1002/cphc.201700782.
[22] S. Li, S. Guo, D. Gong, N. Kang, K.G. Fang, Y. Liu, Nano composite composed of
The authors report no declarations of interest. MoOx -La2O3 –Ni on SiO2 for storing hydrogen into CH4 via CO2 methanation, Int.
J. Hydrogen Energy 44 (2019) 1597–1609, https://doi.org/10.1016/j.
ijhydene.2018.11.130.
Acknowledgements [23] P.A.U. Aldana, F. Ocampo, K. Kobl, B. Louis, F. Thibault-Starzyk, M. Daturi,
P. Bazin, S. Thomas, A.C. Roger, Catalytic CO2 valorization into CH4 on Ni-based
The authors acknowledge CAPES for the financial support to carry ceria-zirconia. Reaction mechanism by operando IR spectroscopy, Catal. Today
215 (2013) 201–207, https://doi.org/10.1016/j.cattod.2013.02.019.
out this work. [24] R. Razzaq, C. Li, M. Usman, K. Suzuki, S. Zhang, A highly active and stable Co4N/
γ-Al2O3 catalyst for CO and CO2 methanation to produce synthetic natural gas
(SNG), Chem. Eng. J. 262 (2015) 1090–1098, https://doi.org/10.1016/j.
Appendix A. Supplementary data
cej.2014.10.073.
[25] X. Guo, A. Traitangwong, M. Hu, C. Zuo, V. Meeyoo, Z. Peng, C. Li, Carbon dioxide
Supplementary material related to this article can be found, in the methanation over nickel-based catalysts supported on various mesoporous
material, Energy Fuels 32 (2018) 3681–3689, https://doi.org/10.1021/acs.
online version, at doi:https://doi.org/10.1016/j.jece.2020.104629.
energyfuels.7b03826.
[26] F. Song, Q. Zhong, Y. Yu, M. Shi, Y. Wu, J. Hu, Y. Song, Obtaining well-dispersed
References Ni/Al2O3 catalyst for CO2 methanation with a microwave-assisted method, Int. J.
Hydrogen Energy 42 (2017) 4174–4183, https://doi.org/10.1016/j.
ijhydene.2016.10.141.
[1] J. Kirchner, J.K. Anolleck, H. Lösch, S. Kureti, Methanation of CO2 on iron based
[27] A.H. Zamani, R. Ali, W.A.W. Abu Bakar, Optimization of CO2 methanation reaction
catalysts, Appl. Catal. B Environ. 223 (2018) 47–59, https://doi.org/10.1016/j.
over M*/Mn/Cu-Al2O3 (M*: Pd, Rh and Ru) catalysts, J. Ind. Eng. Chem. 29 (2015)
apcatb.2017.06.025.
238–248, https://doi.org/10.1016/j.jiec.2015.02.028.
[2] W. Wang, J. Gong, Methanation of carbon dioxide: an overview, Front. Chem. Eng.
[28] P. Panagiotopoulou, Hydrogenation of CO2 over supported noble metal catalysts,
China. 5 (2011) 2–10, https://doi.org/10.1007/s11705-010-0528-3.
Appl. Catal. A Gen. 542 (2017) 63–70, https://doi.org/10.1016/j.
[3] Z.Z. Noor, R.O. Yusuf, A.H. Abba, M.A. Abu Hassan, M.F. Mohd Din, An overview
apcata.2017.05.026.
for energy recovery from municipal solid wastes (MSW) in Malaysia scenario,
[29] D. Wierzbicki, R. Baran, R. Dębek, M. Motak, T. Grzybek, M.E. Gálvez, P. Da Costa,
Renew. Sust. Energy Rev. 20 (2013) 378–384, https://doi.org/10.1016/j.
The influence of nickel content on the performance of hydrotalcite-derived
rser.2012.11.050.
catalysts in CO2 methanation reaction, Int. J. Hydrogen Energy 42 (2017)
[4] K. Ray, G. Deo, A potential descriptor for the CO2 hydrogenation to CH4 over Al2O3
23548–23555, https://doi.org/10.1016/j.ijhydene.2017.02.148.
supported Ni and Ni-based alloy catalysts, Appl. Catal. B Environ. 218 (2017)
[30] B. Mutz, P. Sprenger, W. Wang, D. Wang, W. Kleist, J.D. Grunwaldt, Operando
525–537, https://doi.org/10.1016/j.apcatb.2017.07.009.
Raman spectroscopy on CO2 methanation over alumina-supported Ni, Ni3Fe and
[5] L. Xu, F. Wang, M. Chen, J. Zhang, K. Yuan, L. Wang, K. Wu, G. Xu, W. Chen, CO2
methanation over a Ni based ordered mesoporous catalyst for the production of

7
Y.R. Dias and O.W. Perez-Lopez Journal of Environmental Chemical Engineering 9 (2021) 104629

NiRh0.1 catalysts: role of carbon formation as possible deactivation pathway, Appl. [54] P. Wu, J. Sun, M. Abbas, P. Wang, Y. Chen, J. Chen, Hydrophobic SiO2 supported
Catal. A Gen. 556 (2018) 160–171, https://doi.org/10.1016/j.apcata.2018.01.026. Fe-Ni bimetallic catalyst for the production of high-calorie synthetic natural gas,
[31] C. Lv, L. Xu, M. Chen, Y. Cui, X. Wen, Y. Li, C.E. Wu, B. Yang, Z. Miao, X. Hu, Appl. Catal. A Gen. 590 (2020) 117302, https://doi.org/10.1016/j.
Q. Shou, Recent progresses in constructing the highly efficient Ni based catalysts apcata.2019.117302.
with advanced low-temperature activity toward CO2 methanation, Front. Chem. 8 [55] M. Wu, D.M. Hercules, Studies of supported nickel catalysts by X-ray photoelectron
(2020) 1–32, https://doi.org/10.3389/fchem.2020.00269. and ion scattering spectroscopies, J. Phys. Chem. 83 (1979) 2003–2008, https://
[32] M.A.A. Aziz, A.A. Jalil, S. Triwahyono, A. Ahmad, CO2methanation over doi.org/10.1021/j100478a015.
heterogeneous catalysts: Recent progress and future prospects, Green Chem. 17 [56] Y.R. Dias, O.W. Perez-Lopez, Carbon dioxide methanation over Ni-Cu/SiO2
(2015) 2647–2663, https://doi.org/10.1039/c5gc00119f. catalysts, Energy Convers. Manage. 203 (2020), https://doi.org/10.1016/j.
[33] S. Papari, M. Kazemeini, M. Fattahi, M. Fatahi, Dme direct synthesis from syngas in enconman.2019.112214.
a large-scale three-phase slurry bubble column reactor: transient modeling, Chem. [57] F.M. Berndt, O.W. Perez-Lopez, Catalytic decomposition of methane over Ni/SiO2:
Eng. Commun. 201 (2014) 612–634, https://doi.org/10.1080/ influence of Cu addition, React. Kinet. Mech. Catal. 120 (2017) 181–193, https://
00986445.2013.782292. doi.org/10.1007/s11144-016-1096-4.
[34] H.T.T. Nguyen, Y. Kumabe, S. Ueda, K. Kan, M. Ohtani, K. Kobiro, Highly durable [58] C.O. Calgaro, O.W. Perez-Lopez, Decomposition of methane over Co3− xAlxO4
Ru catalysts supported on CeO2 nanocomposites for CO2 methanation, Appl. Catal. (x=0–2) coprecipitated catalysts: the role of Co phases in the activity and stability,
A Gen. 577 (2019) 35–43, https://doi.org/10.1016/j.apcata.2019.03.011. Int. J. Hydrogen Energy 42 (2017) 29756–29772, https://doi.org/10.1016/j.
[35] H. Takano, H. Shinomiya, K. Izumiya, N. Kumagai, H. Habazaki, K. Hashimoto, ijhydene.2017.10.082.
CO2 methanation of Ni catalysts supported on tetragonal ZrO2 doped with Ca2+ [59] T.K.R. de Oliveira, M. Rosset, O.W. Perez-Lopez, Ethanol dehydration to diethyl
and Ni2+ ions, Int. J. Hydrogen Energy 40 (2015) 8347–8355, https://doi.org/ ether over Cu-Fe/ZSM-5 catalysts, Catal. Commun. 104 (2018) 32–36, https://doi.
10.1016/j.ijhydene.2015.04.128. org/10.1016/j.catcom.2017.10.013.
[36] W. Li, X. Nie, X. Jiang, A. Zhang, F. Ding, M. Liu, Z. Liu, X. Guo, C. Song, ZrO2 [60] F. Denardin, O.W. Perez-Lopez, Tuning the acidity and reducibility of Fe/ZSM-5
support imparts superior activity and stability of Co catalysts for CO2 methanation, catalysts for methane dehydroaromatization, Fuel 236 (2019) 1293–1300, https://
Appl. Catal. B Environ. 220 (2018) 397–408, https://doi.org/10.1016/j. doi.org/10.1016/j.fuel.2018.09.128.
apcatb.2017.08.048. [61] O.W. Perez-Lopez, A. Senger, N.R. Marcilio, M.A. Lansarin, Effect of composition
[37] H. Li, S. Cheng, Y. He, M. Javed, G. Yang, R. Yang, N. Tsubaki, A study on the effect and thermal pretreatment on properties of Ni-Mg-Al catalysts for CO2 reforming of
of pH value of impregnation solution in nickel catalyst preparation for methane dry methane, Appl. Catal. A Gen. 303 (2006) 234–244, https://doi.org/10.1016/j.
reforming reaction, ChemistrySelect. 4 (2019) 8953–8959, https://doi.org/ apcata.2006.02.024.
10.1002/slct.201901910. [62] M. Rosset, O.W. Perez-Lopez, Cu–Ca–Al catalysts derived from hydrocalumite and
[38] P. Munnik, P.E. De Jongh, K.P. De Jong, Recent developments in the synthesis of their application to ethanol dehydrogenation, React. Kinet. Mech. Catal. 126
supported catalysts, Chem. Rev. 115 (2015) 6687–6718, https://doi.org/10.1021/ (2019) 497–511, https://doi.org/10.1007/s11144-018-1513-y.
cr500486u. [63] C. Escobar, O.W. Perez-Lopez, Hydrogen production by methane decomposition
[39] B.A.T. Mehrabadi, S. Eskandari, U. Khan, R.D. White, J.R. Regalbuto, A Review of over Cu-Co-Al mixed oxides activated under reaction conditions, Catal. Letters 144
Preparation Methods for Supported Metal Catalysts, 1st ed., Elsevier Inc., 2017 (2014) 796–804, https://doi.org/10.1007/s10562-014-1234-4.
https://doi.org/10.1016/bs.acat.2017.10.001. [64] D.S. Lima, C.O. Calgaro, O.W. Perez-lopez, Biomass and Bioenergy Hydrogen
[40] T.A. Le, M.S. Kim, S.H. Lee, T.W. Kim, E.D. Park, CO and CO2 methanation over production by glycerol steam reforming over Ni based catalysts prepared by di ff
supported Ni catalysts, Catal. Today 293–294 (2017) 89–96, https://doi.org/ erent methods, Biomass Bioenergy 130 (2019) 105358, https://doi.org/10.1016/j.
10.1016/j.cattod.2016.12.036. biombioe.2019.105358.
[41] Z. Zhang, T. Wei, G. Chen, C. Li, D. Dong, W. Wu, Q. Liu, X. Hu, Understanding [65] N.A. Hermes, M.A. Lansarin, O.W. Perez-Lopez, Catalytic decomposition of
correlation of the interaction between nickel and alumina with the catalytic methane over M-Co-Al catalysts (M = Mg, Ni, Zn, Cu), Catal. Letters 141 (2011)
behaviors in steam reforming and methanation, Fuel. 250 (2019) 176–193, 1018–1025, https://doi.org/10.1007/s10562-011-0611-5.
https://doi.org/10.1016/j.fuel.2019.04.005. [66] C.O. Calgaro, O.W. Perez-Lopez, Biogas dry reforming for hydrogen production
[42] R. Daroughegi, F. Meshkani, M. Rezaei, Enhanced activity of CO2 methanation over over Ni-M-Al catalysts (M = Mg, Li, Ca, La, Cu, Co, Zn), Int. J. Hydrogen Energy 44
mesoporous nanocrystalline Ni–Al2O3 catalysts prepared by ultrasound-assisted co- (2019) 17750–17766, https://doi.org/10.1016/j.ijhydene.2019.05.113.
precipitation method, Int. J. Hydrogen Energy 42 (2017) 15115–15125, https:// [67] B. Mile, D. Stirling, M.A. Zammitt, A. Lovell, M. Webb, The location of nickel oxide
doi.org/10.1016/j.ijhydene.2017.04.244. and nickel in silica-supported catalysts: two forms of “NiO” and the assignment of
[43] S. Valinejad Moghaddam, M. Rezaei, F. Meshkani, R. Daroughegi, Carbon dioxide temperature-programmed reduction profiles, J. Catal. 114 (1988) 217–229,
methanation over Ni-M/Al2O3 (M: Fe, Co, Zr, La and Cu) catalysts synthesized https://doi.org/10.1016/0021-9517(88)90026-7.
using the one-pot sol-gel synthesis method, Int. J. Hydrogen Energy 43 (2018) [68] M. Subrahmanyam, A. Venugopal, V. Durga Kumari, J. Ashok, K.B.S. Prasad,
16522–16533, https://doi.org/10.1016/j.ijhydene.2018.07.013. S. Naveen Kumar, D. Hari Prasad, Hydrogen production by catalytic decomposition
[44] K. Ray, S. Sengupta, G. Deo, Reforming and cracking of CH4 over Al2O3 supported of methane over Ni/SiO2, Int. J. Hydrogen Energy 32 (2007) 1782–1788, https://
Ni, Ni-Fe and Ni-Co catalysts, Fuel Process. Technol. 156 (2017) 195–203, https:// doi.org/10.1016/j.ijhydene.2007.01.007.
doi.org/10.1016/j.fuproc.2016.11.003. [69] M. Thommes, K. Kaneko, A.V. Neimark, J.P. Olivier, F. Rodriguez-Reinoso,
[45] J. Károlyi, M. Németh, C. Evangelisti, G. Sáfrán, Z. Schay, A. Horváth, F. Somodi, J. Rouquerol, K.S.W. Sing, Physisorption of gases, with special reference to the
Carbon dioxide reforming of methane over Ni–In/SiO2 catalyst without coke evaluation of surface area and pore size distribution (IUPAC Technical Report),
formation, J. Ind. Eng. Chem. 58 (2018) 189–201, https://doi.org/10.1016/j. Pure Appl. Chem. 87 (2015) 1051–1069, https://doi.org/10.1515/pac-2014-1117.
jiec.2017.09.024. [70] R. Daroughegi, F. Meshkani, M. Rezaei, Characterization and evaluation of
[46] J. Ren, X. Qin, J.Z. Yang, Z.F. Qin, H.L. Guo, J.Y. Lin, Z. Li, Methanation of carbon mesoporous high surface area promoted Ni-Al2O3 catalysts in CO2 methanation,
dioxide over Ni-M/ZrO2 (M = Fe, Co, Cu) catalysts: effect of addition of a second J. Energy Inst. 93 (2020) 482–495, https://doi.org/10.1016/j.joei.2019.07.003.
metal, Fuel Process. Technol. 137 (2015) 204–211, https://doi.org/10.1016/j. [71] B. Alrafei, I. Polaert, A. Ledoux, F. Azzolina-Jury, Remarkably stable and efficient
fuproc.2015.04.022. Ni and Ni-Co catalysts for CO2 methanation, Catal. Today (2019), https://doi.org/
[47] C. Liang, Z. Ye, D. Dong, S. Zhang, Q. Liu, G. Chen, C. Li, Y. Wang, X. Hu, 10.1016/j.cattod.2019.03.026.
Methanation of CO2: impacts of modifying nickel catalysts with variable-valence [72] T.A. Le, J.K. Kang, E.D. Park, CO and CO2 methanation over Ni/SiC and Ni/SiO2
additives on reaction mechanism, Fuel. 254 (2019) 115654, https://doi.org/ catalysts, Top. Catal. 61 (2018) 1537–1544, https://doi.org/10.1007/s11244-018-
10.1016/j.fuel.2019.115654. 0965-7.
[48] G. Garbarino, C. Wang, T. Cavattoni, E. Finocchio, P. Riani, M. Flytzani- [73] D. Wierzbicki, M.V. Moreno, S. Ognier, M. Motak, T. Grzybek, P. Da Costa, M.
Stephanopoulos, G. Busca, A study of Ni/La-Al2O3 catalysts: a competitive system E. Gálvez, Ni-Fe layered double hydroxide derived catalysts for non-plasma and
for CO2 methanation, Appl. Catal. B Environ. 248 (2019) 286–297, https://doi. DBD plasma-assisted CO2 methanation, Int. J. Hydrogen Energy (2019) 2–11,
org/10.1016/j.apcatb.2018.12.063. https://doi.org/10.1016/j.ijhydene.2019.06.095.
[49] Q. Liu, B. Bian, J. Fan, J. Yang, Cobalt doped Ni based ordered mesoporous [74] M. Feyzi, L. Norouzi, Y. Zamani, Preparation and characterization of Fe–Co/SiO2
catalysts for CO2 methanation with enhanced catalytic performance, Int. J. nanocatalysts for gasoline range hydrocarbons production from syngas, Catal.
Hydrogen Energy 43 (2018) 4893–4901, https://doi.org/10.1016/j. Letters 146 (2016) 1922–1933, https://doi.org/10.1007/s10562-016-1839-x.
ijhydene.2018.01.132. [75] G. Zhou, H. Liu, Y. Xing, S. Xu, H. Xie, K. Xiong, CO2 hydrogenation to methane
[50] L. Xu, X. Lian, M. Chen, Y. Cui, F. Wang, W. Li, B. Huang, CO2 methanation over over mesoporous Co/SiO2 catalysts: effect of structure, J. CO2 Util. 26 (2018)
Co–Ni bimetal-doped ordered mesoporous Al2O3 catalysts with enhanced low- 221–229, https://doi.org/10.1016/j.jcou.2018.04.023.
temperature activities, Int. J. Hydrogen Energy 43 (2018) 17172–17184, https:// [76] X. Sun, F. Du, Synthesis, characterization and catalytic properties of monometal/
doi.org/10.1016/j.ijhydene.2018.07.106. SiO2 and bimetal/SiO2 hollow spheres with mesoporous structure, Nano 12 (2017)
[51] D. Pandey, G. Deo, Promotional effects in alumina and silica supported bimetallic 9–14, https://doi.org/10.1142/S179329201750148X.
Ni-Fe catalysts during CO2 hydrogenation, J. Mol. Catal. A Chem. 382 (2014) [77] T. Schaaf, J. Grünig, M.R. Schuster, T. Rothenfluh, A. Orth, Methanation of CO2 -
23–30, https://doi.org/10.1016/j.molcata.2013.10.022. storage of renewable energy in a gas distribution system, Energy Sustain. Soc. 4
[52] D. Pandey, G. Deo, Effect of support on the catalytic activity of supported Ni-Fe (2014) 1–14, https://doi.org/10.1186/s13705-014-0029-1.
catalysts for the CO2 methanation reaction, J. Ind. Eng. Chem. 33 (2016) 99–107, [78] S.S. Kim, H.H. Lee, S.C. Hong, A study on the effect of support’s reducibility on the
https://doi.org/10.1016/j.jiec.2015.09.019. reverse water-gas shift reaction over Pt catalysts, Appl. Catal. A Gen. 423–424
[53] M. Guo, G. Lu, The difference of roles of alkaline-earth metal oxides on silica- (2012) 100–107, https://doi.org/10.1016/j.apcata.2012.02.021.
supported nickel catalysts for CO2 methanation, RSC Adv. 4 (2014) 58171–58177, [79] R. Daroughegi, F. Meshkani, M. Rezaei, Characterization and evaluation of
https://doi.org/10.1039/c4ra06202g. mesoporous high surface area promoted Ni- Al2O3 catalysts in CO2 methanation,
J. Energy Inst. (2019), https://doi.org/10.1016/j.joei.2019.07.003.

You might also like