You are on page 1of 9

Applied Catalysis B: Environmental 270 (2020) 118859

Contents lists available at ScienceDirect

Applied Catalysis B: Environmental


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

Dry reforming of methane over the cobalt catalyst: Theoretical insights into T
the reaction kinetics and mechanism for catalyst deactivation
Shuyue Chena,b,c, Jeremie Zaffrana, Bo Yanga,*
a
School of Physical Science and Technology, Shanghai Tech University, Shanghai 201210, China
b
CAS Key Laboratory of Low-Carbon Conversion Science and Engineering, Shanghai Advanced Research Institute, Chinese Academy of Sciences, Shanghai, 201210, China
c
University of Chinese Academy of Sciences, Beijing, 100049, China

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

Keywords: Cobalt shows high catalytic activity for the important dry reforming of methane (DRM) reaction. However, it is
Dry reforming of methane prone to deactivation and the corresponding mechanism remains controversial. In this work, we combined
Cobalt catalyst density functional theory calculations and microkinetic modeling to study the active site and reaction me-
Catalyst deactivation chanism of Co catalyzed DRM reaction, employing face centered cubic Co(111) and Co(211) as models. It was
Reaction kinetics
found that the step site over Co(211) is the active site for the reaction, and on Co(111), the C + O and CH + O
DFT
paths are the preferred reaction pathways, while the C + O path is dominant on Co(211). The dissociation of
CH4 is the rate-controlling step of DRM over both Co(111) and Co(211). We found that Co(111) is mainly
deactivated due to carbon deposition whilst Co(211) undergoes surface oxidization. In addition, Co(111) tends
to follow the surface carbon coupling mechanism, and surface carbon clusters formed will lead to catalyst de-
activation.

1. Introduction catalyst is cobalt [16,17]. Previous studies indicated that Co could show
a nickel-like activity for the DRM reaction, but the stability problems
According to the global greenhouse gas emissions data released by still exists, and the deactivation mechanism of Co is still under debate
Intergovernmental Panel on Climate Change (IPCC) in 2014, carbon [18–22].
dioxide accounts for 65 % of total greenhouse gas emissions, followed In comparison to the detailed analysis on the deactivation me-
by methane. The greenhouse effect caused by greenhouse gas emissions chanism of Co catalysts in Fischer-Tropsch synthesis studies reported in
can have a great impact on the ecological environment. In addition to the literature [23–25], similar discussions from a computational per-
emissions reduction, developing a process to utilize greenhouse gases spective under DRM reaction conditions is missing. Experimental evi-
provides a new direction for solving this problem. Methane reforming dence showed that, in the DRM process, the Co catalyst can be deac-
with carbon dioxide, i.e. dry reforming of methane (DRM), converts tivated mainly through two mechanisms, namely oxidation and carbon
these two greenhouse gases into two clean fuel gases, carbon monoxide deposition [21,26–29]. It was demonstrated experimentally that the
and hydrogen, i.e. synthetic gas (syngas) [1,2], which can be further deactivation mechanism is strongly dependent on the loading of Co
used in the Fischer-Tropsch process to synthesize desired hydrocarbons [21,26–29]. While highly metal-loaded catalysts are mainly deactivated
[3,4]. However, due to the lack of highly active and stable catalysts, the via coke formation, low metal loading mainly leads to Co oxidation
attractive DRM reaction is still not maturely industrialized. [26–28]. Jose-Alonso et al. found that the 1 wt % Co catalyst was
It was found that highly active and deactivation-resistant precious abruptly deactivated by oxidation, while the catalyst with a Co content
metal catalysts, such as Rh and Ru, can be used for DRM [5–7], but the of 9 wt % showed high carbon deposition due to the large size of Co
high price and low reserves limit their further development. In addition particles present on the support [21]. Therefore, the deactivation might
to the noble metals, cheap and earth-abundant transition metals are be related to the size of metallic nanoparticles. It could be further in-
considered as alternatives [8]. Ni has attracted considerable attentions ferred that differences in size may result in different ratios of surface
due to its high activity and large reserves, but the disadvantage of the active sites, resulting in different deactivation mechanisms.
Ni catalyst is also obvious that it is easily deactivated through carbon In general, the support can affect the size and structure of the metal
deposition [9–15]. Another widely used inexpensive transition metal particle [30,31], form interfacial active sites with the metal [32], and


Corresponding author.
E-mail address: yangbo1@shanghaitech.edu.cn (B. Yang).

https://doi.org/10.1016/j.apcatb.2020.118859
Received 19 January 2020; Received in revised form 3 March 2020; Accepted 5 March 2020
Available online 06 March 2020
0926-3373/ © 2020 Elsevier B.V. All rights reserved.
S. Chen, et al. Applied Catalysis B: Environmental 270 (2020) 118859

even directly participate in the catalytic reaction, thereby affecting the and transition states, and the k-point was set to 3 × 3 × 1 and
activity and stability [33]. However, it was found that the presence of 4 × 3 × 1, respectively [61]. The bottom two and six layers in the slabs
support influences the activity and stability mainly by varying the ad- of Co(111) and Co(211), respectively, were fixed at the bulk structure
sorption of intermediate and the energy barriers of key elementary and the rest of the atoms in the slabs were allowed to relax during the
steps in the reaction. Here, we focus on how the activity and stability of structural optimizations. The vacuum space introduced for each slab
Co in DRM reactions are affected by species adsorption and activation was with a height of 15 Å to avoid the periodic interactions. The ad-
barriers [34]. These insights should also be applicable to the discussion sorption energy of surface species was calculated using the following
of the activity and stability of supported Co catalysts. equation.
We noticed that there are two crystallographic structures of Co, i.e. Ead = Eslab/ads − Eads − Eslab (1)
face centered cubic (FCC) and hexagon close-packed (HCP), and tran-
sition between these two structures can happen under different condi- Where, Ead is the adsorption energy, Eslab/ads, Eads and Eslab is the energy
tions [35–38]. Typically, at low temperatures, Co prefers the form of of slab with adsorbates, the energy of adsorbate in gas phase and the
HCP phase, and when the temperature rises to 700 K and above, HCP Co energy of slab, respectively.
will be transformed to FCC Co [35]. Kitakami and co-workers also It was found by several studies that the calculations of gaseous CO2
studied the effect of Co particle size on the crystalline phase below and CO may be inaccurate using DFT [62,63]. One possible approach to
673 K. It was found that when the metal particle diameter is smaller solve this problem is to correct gaseous energies of CO2 or CO to make
than 200 Å, cobalt is in the form of FCC phase, and when the particle sure that the calculated reaction Gibbs free energies with DFT data are
diameter is larger than 400 Å, cobalt would prefer the HCP crystal identical to the experimental values. In this work, the standard reaction
structure [37]. Considering that the typical operating temperature of Gibbs free energy calculated for the DRM reaction at 298 K is 1.90 eV,
the DRM reaction is around 873∼1123 K and the diameter of Co par- which is very close to the experimental value of 1.77 eV, suggesting that
ticles are usually less than 400 Å [39]. It was also found experimentally no correction should be made for the energies of CO2 and CO here.
that under the conditions of DRM reaction, Co exists as FCC structure on
different oxide supports [40,41]. Therefore, we adopt the FCC Co in the 2.2. Microkinetic modeling
current study. Since the deactivation process may be structure sensitive,
we also investigate the reactions on different facets of FCC Co. It was Steady state approximation and transition state theory were used for
widely reported in the literature that smaller particles would possess microkinetic simulation. The CatMAP software package developed by
more low-coordination step sites such as fcc-(211), whilst larger par- Nørskov’s group was used [64], which has been proved to be a powerful
ticles show a higher ratio of fcc-(111) terrace sites [42,43]. It was also tool for obtaining kinetic insights into surface catalytic reactions by
found experimentally and theoretically that Co(111) is the dominant applying the mean-field assumption [49–54,65,66]. The simulation
surface of large FCC Co nanoparticles [41], and on small-sized particles temperature was set to 873∼1073 K and the total pressure was 1 bar.
stepped surfaces are abundant which may give rise to a different de- Methane conversion considered was 5% and the ratio between CH4 and
activation mechanism. CO2 in the gaseous reactant was 1:1. Further details of the thermo-
To the best of our knowledge, there is no thorough computational dynamic corrections for gaseous species and surface intermediates were
study in the literature reporting the reaction kinetics of DRM at dif- given in the Supporting Information (SI).
ferent sites over Co and a detailed analysis on the catalyst deactivation We also resorted to the degree of rate control (DRC) analysis method
process. In this work, we used the (111) and (211) surfaces of FCC Co as developed by the Campbell group [67,68], which has been proved as a
models to represent different surface sites on the catalyst. Using DFT useful tool to probe the surface key intermediates and transition states
calculations combined with microkinetic simulations, reaction rates of in the catalytic process [49–54,65,69,70]. The values of DRC can be
possible reaction pathways of DRM on different Co surfaces were cal- determined by calculating the response of reaction rate to changes in
culated, and the corresponding surface dominant species and rate the free energy of the intermediate and transition states using the fol-
controlling species were analyzed. Finally, insights into the mechanism lowing equation:
of deactivation of Co catalysts during the DRM reactions on different
surfaces were provided, and possible strategies for future design of −∂ ln r ⎞
Xi = ⎜⎛ 0 ⎟
active and stable Co catalysts were suggested. ⎝ Gi / kB T ) ⎠Gj0≠ i
∂ (
(2)

2. Computational methods Where, r is the overall reaction rate and G is the free energy of each
intermediate or transition state, kB is the Boltzmann constant and T is
2.1. DFT calculations the temperature. For a transition state, the value of DRC should be
between 0 and 1. The higher the DRC value, the greater the effect of the
All the DFT calculations were performed with the VASP software transition state on the rate. The value of intermediates should be 0 or
[44–46]. The projector-augmented wave (PAW) method was used to negative and related to the coverage of the intermediate and the
describe the ion core and its interaction with valence electrons [47]. number of site required by the rate controlling transition state (RCTS)
Electron exchange and correlation was treated with the BEEF-vdW [67,68].
functional, which is widely used in surface properties calculations
[48–55]. The energy cutoff was set to 500 eV. The energy convergence 3. Results and discussion
criterion was set to 10−4 eV, and the force convergence was set to
0.05 eV/Å. We also included the spin polarization in all the calcula- 3.1. Reaction network of the DRM reaction on Co surfaces
tions, and transition states were searched using the constrained mini-
mization method [56–58]. Scheme 1 shows the possible DRM reaction network, including three
We first optimized the lattice of FCC Co. The optimized lattice CO2 activation steps and different pathways from methane to the pro-
parameter was 3.536 Å, which was found close to the experimental duct CO, which are similar to those reported in the literature on Ni, Pt
value of 3.550 Å. And the magnetic moment of Co calculated in this and Pd catalysts [71–73]. We divide the elementary steps into the fol-
work is 1.72 μB/atom, which is very close to the experimental value lowing types: (1) Direct and H-assisted dissociation of CO2; (2) Hy-
(1.71 μB/atom) and previous theoretical results (1.63 μB/atom) drogenation of O* to OH*; (3) Dehydrogenation of CH4 to CHx
[59,60]. Then we built four-layer p(3 × 3) Co(111) and twelve-layer p (x = 0–3); (4) CHx* coupling with O*/OH* to form CHxO*/CHxOH*;
(1 × 3) Co(211) slabs for the calculations related with the adsorption (5) Dehydrogenation of CHxO*/CHxOH* and (6) The coupling of

2
S. Chen, et al. Applied Catalysis B: Environmental 270 (2020) 118859

Table 1
Reaction energy (ΔE) and activation barrier (Ea) of the DRM elementary steps
on Co(111) and Co(211).
No. elementary steps Co(111)  Co(211)

ΔE/eV Ea/eV ΔE/eV Ea/eV

1 CO2 + 2* → CO* + O* −0.90 0.50 −1.25 0.09


2 CO2 + H* → COOH* 0.24 1.36 −0.14 1.40
3 COOH* + * →CO* + OH* −1.10 0.58 −1.24 0.60
4 CO2 + H* → HCOO* −0.40 0.82 −1.11 0.73
5 HCOO*+ * → CHO* + O* 0.58 1.12 0.51 2.45
6 CH4 + 2* → CH3* + H* 0.25 1.17 0.09 0.84
7 CH3* + * → CH2* + H* 0.36 0.89 0.58 0.79
8 CH2* + * → CH* + H* −0.23 0.37 −0.32 0.36
9 CH* + * → C* + H* 0.39 1.23 0.05 0.77
10 CH3* + O* → CH3O* + * 0.08 1.56 0.36 2.00
11 CH3O* + * → CH2O* + H* 0.71 0.34 0.75 0.95
12 CH3* + OH* → CH3OH* + * 0.65 2.12 1.20 2.38
13 CH3OH* + * → CH3O* + H* −0.62 1.02 −0.71 0.90
14 CH3OH* + * → CH2OH* + H* 0.50 1.22 0.29 1.05
15 CH2* + O* → CH2O* + * 0.43 1.24 0.53 1.58
16 CH2O* + * → CHO* + H* −0.08 0.34 −0.02 0.45
17 CH2* + OH* → CH2OH* + * 0.79 1.49 0.91 1.48
18 CH2OH* + * → CH2O* + H* −0.40 0.80 −0.25 0.94
19 CH2OH* + * → CHOH* + H* −0.03 0.63 0.17 0.48
20 CH* + O* → CHO* + * 0.58 1.29 0.83 1.67
21 CHO* + * → CO* + H* −1.09 0.16 −0.65 0.05
22 CH* + OH* → CHOH* + * 0.99 1.62 1.41 2.19
23 CHOH* + * → H* + COH* −0.61 0.12 −0.21 0.03
24 CHOH* + * → H* + CHO* −0.46 0.90 −0.44 1.30
25 C* + O* → CO* + * −0.90 1.57 0.13 1.75
26 C* + OH*→ COH* + * −0.01 1.61 1.15 1.79
27 COH* + * → CO* + H* −0.93 0.32 −0.88 0.63
28 O* + H* → OH* + * 0.04 1.22 −0.13 1.56
29 CO* → CO + * 1.46 – 1.46 –
30 H* + H* → H2 + 2* 0.75 – 0.59 –
Scheme 1. (a) Three pathways of carbon dioxide activation; (b) Reaction
scheme from methane to CO, the blue (red) arrows indicate the reactions be-
tween CHx (x = 1–3) and OH* (O*), whilst the black arrows represent dehy- to break the CeH bond; (4) The barrier of CHx* coupling with O* or
drogenation reactions. The dash and solid lines indicate the breaking of OeH
OH* on Co(211) is usually higher than on Co(111). However, it is im-
and CeH bonds, respectively.
practical to determine the reactivity and main reaction pathways only
from the barriers and reaction energies listed in Table 1. Therefore, we
surface H* to form gaseous H2. The configurations of adsorption and further carried out microkinetic simulations over Co(111) and Co(211).
transition states are shown in Figs. S1–S3 of the SI.
3.2. Microkinetic analysis of the DRM reaction on Co surfaces
3.1.1. CO2 activation steps
Scheme 1(a) shows different CO2 activation pathways for CO pro- 3.2.1. Reaction mechanism and active sites
duction. One can see that CO2 can be directly decomposed into CO* and In the microkinetic simulations, all the elementary steps presented
O*, or combined with the H* produced from methane dehydrogenation, in Table 1 are taken into consideration, and the reaction rates of the
to give COOH* or HCOO*. COOH* is then decomposed into either CO* elementary steps at 973 K are listed in Table 2. After analyzing the
and OH* or COH* and O* followed by COH* dehydrogenation to reaction rates, the reactivity of the two Co surfaces and the main DRM
generate CO. Dissociation of HCOO* would produce O* and CHO* reaction pathway can be determined. The reaction rate of the dominant
followed by CHO* dehydrogenation to generate CO. The reaction bar- reaction pathways of DRM are plotted in Fig. 1 as a function of tem-
riers of different elementary steps in the pathways of CO2 activation are perature. It can be found from this figure that the reaction rate on Co
listed in Table 1. One can see from Table 1 that the barrier of CO2 direct (111) is lower than that on Co(211) at the temperatures considered
dissociation (R1) is much lower than that of H* coupling with CO2 (R2 here. In addition, the reaction rate on Co(111) is almost identical to that
and R3) on both Co(111) and Co(211), indicating that the direct dis- obtained by our group on Ni(111) [71]. The difference in the reaction
sociation of CO2 might be favored in comparison with the H-assisted rate of two cobalt surfaces decreases from about 1.5 to 0.5 orders of
activation pathways. magnitude when the temperature increases from 873 K to 1073 K. This
indicates that the step sites should be the active site at low reaction
3.1.2. From CH4 to CO temperatures, but the reactivity of the terrace and step sites becomes
Scheme 1(b) shows the different reaction pathways from CH4 to H2 identical at high reaction temperatures.
and CO. Methane can be dehydrogenated to produce CHx (x = 0∼3) Regarding the mechanism of CO2 activation, one can see from
species. Then the CHx species can couple with O * or OH * to form CHxO Table 2 that the rate of CO2 direct dissociation is at least five orders of
or CHxOH. These surface intermediates will undergo further dehy- magnitude higher than that of H-assisted dissociation on both Co(111)
drogenation until the final product CO is formed. The energetics listed and Co(211), indicating that the direct dissociation of CO2 is favored.
Table 1 for the two Co surfaces reflect the following trends: (1) For This is the same with the way of carbon dioxide activation on Ni,
CHx*, the dehydrogenation barrier is lower than the corresponding whereas Pd and Pt favor the hydrogen-assisted carbon dioxide dis-
coupling barrier of with O*/OH*; (2) Compared with CHx* coupling sociation [71,73,74]. The dominant reaction path on Co(111) is also the
with OH*, the association with O* is kinetically favored; (3) For same as that on Ni(111) [71]. On Co(111), there are two dominant
CHxOH* species, the barrier for the OeH bond breaking is higher than reaction pathways for CO formation, namely the C + O path and

3
S. Chen, et al. Applied Catalysis B: Environmental 270 (2020) 118859

Table 2 The rates of other reaction paths are all at least 3 orders of magnitude
The net reaction rate of each elementary step calculated at the steady state and lower than the rate of the C + O path. In contrast, as shown in Fig. 1(b),
973 K on Co(111) and Co(211). the C + O path can be considered as the dominant reaction path on Co
No. elementary steps reaction rate/site−1 s−1 (211). The rates of other paths are at least 1.5 orders of magnitude
lower than those of the C + O path. Based on our previous work, we
Co(111) Co(211) found that the main reaction paths on Co are the same as on the Ni
surfaces. [75] However, the reaction rate of Ni catalyst is higher, which
1 CO2 + 2* → CO* + O* 9.58 × 10° 4.16 × 101
2 CO2 + H* → COOH* 3.09 × 10−5 5.02 × 10−5 is consistent with the experimental observation [76,77]. Meanwhile, as
3 COOH* + * →CO* + OH* 3.09 × 10−5 5.02 × 10−5 we mentioned in the introduction, the deactivation performance of Ni
4 CO2 + H* → HCOO* 6.27 × 10−5 7.48 × 10−9 and Co is different, which will be discussed in the following sections.
5 HCOO*+ * → CHO* + O* 6.27 × 10−5 7.48 × 10−9
6 CH4 + 2* → CH3* + H* 9.58 × 10° 4.16 × 101
7 CH3* + * → CH2* + H* 9.58 × 10° 4.16 × 101 3.2.2. Coverage of surface species and DRC analysis
8 CH2* + * → CH* + H* 9.58 × 10° 4.16 × 101
We plotted in Fig. 2 the steady state coverage of dominant surface
9 CH* + * → C* + H* 1.05 × 10° 4.14 × 101
10 CH3* + O* → CH3O* + * 4.92 × 10−6 3.03 × 10−6 intermediates, including CO*, C* and O* on two Co surfaces. The
11 CH3O* + * → CH2O* + H* 4.91 × 10−6 3.04 × 10−6 coverage of other surface intermediates is close to 0 and thus omitted
12 CH3* + OH* → CH3OH* + * 1.13 × 10−8 4.80 × 10−1 from Fig. 2. On Co(111), the coverage of CO* is around 0.2 monolayer
13 CH3OH* + * → CH3O* + H* −1.03 × 10-8 5.69 × 10−10
(ML) at 873 K, and decreases to 0 when the temperature is higher than
14 CH3OH* + * → CH2OH* + H* 1.13 × 10−8 1.02 × 10−8
15 CH2* + O* → CH3O* + * −4.44 × 10-6 −3.00 × 10-9
973 K. The coverage of C* on the surface stays at 0 ML at different
16 CH2O* + * → CHO* + H* 5.03 × 10−205 3.80 × 10−118 temperatures, whilst the coverage of O* slightly increases from 0 to
17 CH2* + OH* → CH2OH* + * 6.40 × 10−7 4.40 × 10−7 0.1 ML when the temperature rises from 873 to 1073 K. On Co(211), the
18 CH2OH* + * → CH2O* + H* −4.65 × 10-7 −1.30 × 10-8 coverage of O* increases from 0.6 ML (873 K) to about 0.8 ML (1073 K)
19 CH2OH* + * → CHOH* + H* 1.12 × 10−6 4.53 × 10−7
with temperature. Therefore, one can find that the high coverage of
20 CH* + O* → CHO* + * 8.53 × 10° 2.00 × 10−1
21 CHO* + * → CO* + H* 8.53 × 10° 2.00 × 10−1 oxygen on Co(211) may facilitate the oxidation of Co. The coverage of
22 CH* + OH* → CHOH* + * 7.37 × 10−4 4.11 × 10−6 both CO* and C* on Co(211) is low at 873 K (about 0.1 ML) and drops
23 CHOH* + * → H* + COH* 7.36 × 10−4 4.37 × 10−6 to about 0 ML when the temperature rises.
24 CHOH* + * → H* + CHO* 2.59 × 10−6 1.91 × 10−4
Since the reverse water gas shift (RWGS) reaction, i.e. CO2 + H2 →
25 C* + O* → CO* + * 1.04 × 10° 4.09 × 101
26 C* + OH*→ COH* + * 6.51 × 10−3 4.80 × 10−1
CO + H2O, was reported as a major side reaction in the DRM reaction
27 COH* + * → CO* + H* 7.25 × 10−3 4.80 × 10−1 [78,79], we investigated the effect of this reaction on the coverage of
28 O* + H* → OH* + * 7.22 × 10−3 4.80 × 10−1 O* over Co(111) and Co(211). We added the elementary step of H2O
29 CO* → CO + * 1.92 × 101 8.32 × 101 formation to the kinetic model, with other simulation conditions un-
30 H* + H* → H2 + 2* 1.92 × 101 8.32 × 101
changed. The obtained coverage of O* at steady state are shown in Fig.
S6. One can find from the figure that, upon considering the RWGS re-
CH + O path, as shown in Fig. 1. In the C + O Path, CH4 is dehy- action, the coverage of O* is almost unchanged on Co(111), while on Co
drogenated to C* sequentially, and then C* couples with O*, formed (211), the coverage is reduced by around 0.1 ML. However, the cov-
from CO2 dissociation, to produce CO*. In the CH + O path, CH* erage of O*on Co(211) is still much higher than that on Co(111), even
couples with O* to form CHO* and finally CO* is obtained through with the RWGS reaction included in the modeling.
dehydrogenation. As one can see from Fig. 1(a), on Co(111), the rate of The DRC analysis results are also shown in Fig. 2. One can find from
CH + O path is around 1.5 orders of magnitude higher than that of the the figure that, on both Co(111) and Co(211), transition state of CH4
C + O path at 873 K, and the difference is narrowed down at 1073 K. dissociation is the RCTS at all the temperatures considered. On Co(111),
the DRC value of CO* is around -0.5 at 873 K, and gets close to zero

Fig. 1. Reaction rates of the dominant reaction pathways (highlighted in the above panel) and the total rates of the DRM reaction on Co(111) and Co(211) as a
function of temperature.

4
S. Chen, et al. Applied Catalysis B: Environmental 270 (2020) 118859

Fig. 2. Coverage of surface intermediates and the calculated DRC values of transition states and surface intermediates on Co(111) and Co(211), as a function of
temperatures.

when temperature rises due to the decrease of CO coverage on the E Nads /slab − Eslab − Nμads
ΔG (T ,p,Nads) =
surface. In contrast, the DRC value of O* is close to 0 at 873 K, and N (3)
becomes more negative to -0.25 at 1073 K. On Co(211) surface, O* is
shown as the only rate controlling intermediates and becomes more Where, ΔG (T , p , Nads) , hereinafter referred to as ΔGC* or ΔGO*, is the
rate-controlling with the increase of temperature. Although oxygen is adsorption free energy of C* or O*, N is the number of adsorbates,
not directly participating in the dissociation of CH4, the presence of ENads/slab and Eslab is the energy of slab with N number of adsorbates and
oxygen on the surface will influence the surface free sites that can be the energy of clean surface, respectively. μads is the chemical potential
used for the dissociation reaction where one more free site is needed. of referenced C or O, which can be calculated from μC = μCH4 - 2μH2 and
μO = μCO2(gas) - μCO(gas), respectively. The chemical potential can be
calculated from the DFT energy with thermodynamic corrections at
3.3. Deactivation mechanism of Co 973 K, and the pressures of gaseous species is the same with those used
in the microkinetic simulations.
3.3.1. Stability of C* and O* on two Co surfaces We summarize in Fig. 3 the average adsorption free energy of C*
In order to understand the behavior of C* and O* during the de- and O* on Co(111) and Co(211) as a function of coverage. The surface/
activation of Co catalysts, we hereby calculate the adsorption of C* and subsurface ΔGC* will determine whether the adsorbed carbon can dif-
O* at different coverages on Co(111) and Co(211). For the extra cal- fuse from the surface sites to subsurface and which carbon deposition
culations, we constructed four-layer p(2 × 2) Co(111) and twelve-layer mechanism to follow on different Co surfaces [81]. We find that on Co
p(1 × 4) Co(211) slabs and placed different numbers of adsorbates at (111), the surface/subsurface ΔGC* at 0.25 ML are almost identical,
these sites. The k-points used for the above two slabs were (3 × 3 × 1) indicating that the diffusion of C* between the surface and subsurface
and (4 × 2×1) for Co(111) and Co(211), respectively. Other para- sites is thermodynamically neutral. As the coverage increases, the dif-
meters and convergence criteria are the same with all the previous ference between ΔGC* at the subsurface and surface becomes larger, and
calculations. the surface carbon becomes more stable, indicating that the diffusion of
Typically, carbon deposition can follow the dissolving mechanism C* from surface to subsurface becomes thermodynamically prohibited.
or the surface coupling mechanism. In the dissolving mechanism, the In addition, we find that when the coverage is higher than 0.25 ML,
surface C* species will dissolve into the bulk of metals to form carbide, surface carbon atoms tend to form clusters on Co(111) (see Fig. 3a), and
and then precipitates at the rear side of the nanoparticles to form a are therefore stabilized. However, such cluster structures are not ob-
carbon nanofilament [80]. In the surface coupling mechanism, the served on Co(211), where C* adsorption is destabilized with coverages
surface carbon species will couple with each other and form stable increased. More importantly, surface C* is more stable than subsurface
surface clustered structures [80]. In this section we will consider both C* at any coverage on Co(211), indicating that the diffusion of C* to
mechanisms at different Co sites by determining the behavior of carbon subsurface is also thermodynamically prohibited.
at the surface and subsurface sites with different coverages. Regarding For surface O*, it is also shown in Fig. 3 that higher coverage will
the oxidation process, as mentioned above, the coverage of O* at the result in weaker adsorption on the two cobalt surfaces studied. More
step sites over Co(211) are much higher than that on Co(111), we importantly, the adsorption of O* on Co(211) is always more stable
therefore further verify the coverage dependent adsorption of O* at the than that on Co(111) at all the coverages considered here. Combining
surface sites of both surfaces. with the microkinetic simulation results obtained from low coverage
The average adsorption Gibbs free energy of C* or O* adsorbed at DFT energies shown in Fig. 2, these results suggest that O* deposition
the surface or subsurface site at different coverage can be calculated will preferentially take place on Co(211). Furthermore, the dissociation
using the following equation: barrier of CO2 on Co(211) is lower than on Co(111), indicating that the
produce of O* on Co(211) surface is also kinetically favored. The

5
S. Chen, et al. Applied Catalysis B: Environmental 270 (2020) 118859

and the results are summarized in Fig. 4.


On Co(111), the barrier of CeC coupling (1.07 eV) is lower than that
of the diffusion of C* from surface to subsurface (1.23 eV) and CO*
formation (1.59 eV), indicating the surface coupling mechanism on this
surface is kinetically favored. On Co(211), the barrier of C* and O*
association (1.74 eV) is the lowest among all three elementary steps
considered, indicating that CO formation might be kinetically favored
on this surface. The above results suggest that carbon deposition is
preferred on Co(111), which is consistent with the experimental results
that catalyst with large Co particles is readily deactivated due to carbon
deposition [26,82,83]. Furthermore, we find that the carbon deposition
over cobalt in DRM reaction follows the surface coupling mechanism
rather than the dissolving mechanism.

3.3.3. Competition between the deposition and removal of surface carbon


According to the results presented above, the surface coupling me-
chanism is kinetically favored but thermodynamically unfavored in
comparison with CO formation on Co(111) (Fig. 4a). To further un-
derstand the competition between the deposition and removal of sur-
face carbon on Co(111), we calculated and compared the corresponding
reaction rates at different temperatures. The relevant elementary steps
considered are C* + C* → C2* + * and C* + O* ↔ CO* + *. It should
be noted that the single direction arrow indicates an irreversible reac-
tion while the double direction arrow for a reversible reaction.
We here consider the CeC bonding reaction as an irreversible re-
action for the following reasons. Firstly, under the temperature condi-
tions we considered in this work, the formation of C2* is always an
exergonic process with the reaction Gibbs free energy closed to
−0.64 eV (see Fig. S7). Secondly, C2* is not the end point of surface C*
coupling due to the thermodynamically favored carbon cluster forma-
tion. Finally, the coverage of surface C2* at the beginning of the surface
carbon deposition process is negligible, meaning that the reverse rate of
Fig. 3. Average adsorption free energy (973 K) of carbon and oxygen atoms at the CeC bonding reaction should be low enough to be ignored.
different coverages on (a) Co(111) and (b) Co(211). Here sur and sub are the
Regarding CO* formation, the reverse reaction could not be ignored
surface and subsurface sites, respectively. The average adsorption free energies
since CO* is the main product and the concentration of CO* is much
are calculated with Eq. (3). Figures inserted are the corresponding optimized
configurations of carbon adsorbed at the surface sites.
higher than that of C2*. Therefore, the ratio between the reaction rates
can be calculated from the following equation
+
surface oxidation process of a catalyst is rather complex and may in- rC2 kC-CθC2
= + *−
clude several steps, such as deposition of oxygen, oxygen permeation rCO kC-OθC*θO* − kC-O θCO *θ* (4)
into the bulk and surface reconstruction induced by oxygen. Among all +
Where, kC-C represents the forward rate constant of C2* formation re-
these steps, deposition of surface oxygen should be the first, and higher + −
action, kC-O and kC-O are the forward and reverse rate constant of C* and
coverage of oxygen deposition would indicate that the surface might be
O* association reaction, respectively. θC*, θO*, θCO* and θ* is the cov-
easier to be oxidized
erage of C*, O*, CO* and free site from the above microkinetic simu-
lations, respectively, implying that we are taking the steady state of
3.3.2. Kinetics of carbon deposition DRM over the clean Co surface as the starting point of deactivation. We
After discussing the thermodynamic properties of C* at the surface plot the calculation results of Eq. (4) against the reaction temperature
or subsurface sites on two Co facets, the kinetics of carbon deposition in Fig. 5.
on the surface of p(3 × 3) Co(111) and p(1 × 4) Co(211) are further From Fig. 5, one can see that at 873 K, the rate of C2* formation is
studied. We calculated the barriers of the first step of surface coupling five times faster than that of CO formation. With the increase of tem-
mechanism (CeC coupling), the first step of carbon dissolve mechanism perature, the ratio decreases gradually, and is equal to 1 at the tem-
(C* diffusion to subsurface) and CO* formation on two cobalt surfaces, perature around 973 K. This indicates that below 973 K, the rate of CeC

Fig. 4. Energy profiles of carbon oxidation to


CO (black), carbon-carbon coupling (blue) and
diffusion of surface carbon into subsurface
(red) on Co(111) and Co(211). C** in the fig-
ures represents subsurface carbon. (For inter-
pretation of the references to colour in this
figure legend, the reader is referred to the web
version of this article).

6
S. Chen, et al. Applied Catalysis B: Environmental 270 (2020) 118859

path is the only dominant path on Co(211). It was also found that the
rate-controlling step of the DRM reaction is the same over Co(111) and
Co(211), i.e. the dissociation of CH4. We further studied the deactiva-
tion of two Co surfaces and showed the deactivation mechanism is facet
dependent. Compared with Co(111), the stronger O* adsorption and
lower CO2 dissociation barrier on Co(211) leads to much higher O*
coverage on the surface, indicating that O* deposition will pre-
ferentially take place on Co(211). In contrast, coupling of surface
carbon atoms to carbon clusters is thermodynamically and kinetically
favorable on Co(111). We also highlighted that Co deactivation through
carbon deposition is temperature dependent, and the resistance to
carbon deposition would be improved at high temperatures. Our results
provide insights into understanding experimental observations of the
different deactivation behavior of catalysts with different Co particle
size and at different temperatures, based on which strategies for the
design of Co-based catalysts were proposed.

Fig. 5. Ratio between the rates of carbon-carbon coupling and carbon oxidation CRediT authorship contribution statement
to CO on Co (111) surface, which is calculated with Eq. (4), as a function of
temperature. Higher carbon-carbon coupling rate indicates faster carbon de-
Shuyue Chen: Investigation, Data curation, Formal analysis,
position onto the catalyst surface.
Writing - original draft. Jeremie Zaffran: Writing - review & editing.
Bo Yang: Conceptualization, Supervision, Funding acquisition, Project
coupling is higher than that of CO formation on Co(111). However, at administration, Resources, Writing - review & editing.
high temperatures, i.e. above 973 K, CeC coupling rate tends to de-
crease in comparison with CO formation rate. Hence, our results suggest Declaration of Competing Interest
that carbon deposition is the main mechanism of deactivation on Co
(111) surface at low temperatures, and that high temperatures prevent The authors declare that they have no known competing financial
catalyst deactivation. Such observations agree well with experiments. interests or personal relationships that could have appeared to influ-
Indeed, the coke formation on 10 wt % Co/γ-Al2O3 catalyst is 0.25 g ence the work reported in this paper.
(carbon) g(cat)−1 h−1 at 973 K [84], or 35 % of reacted methane turn
into carbon deposition on 7.6 wt% Co catalyst [85]. At high tempera- Acknowledgements
tures, e.g. 1023 K, no carbon deposition and deactivation during DRM
reaction at 1073 K with 10 wt% Co/Al2O3 catalyst [86]. This work is financially supported by the National Natural Science
Foundation of China (91745102, 21950410522). We thank the HPC
3.4. Insights into future cobalt-based catalyst design Platform of ShanghaiTech University and Shanghai Supercomputer
Center for computing time.
Now we have clarified the reactivity and deactivation mechanism of
the Co catalyst towards the DRM reaction. In general, the trend of both Appendix A. Supplementary data
activity and stability of catalysts with different sizes can be derived
from the ratios of the (211) and (111) surfaces. Smaller Co particles Supplementary material related to this article can be found, in the
showing more low coordination sites would give higher DRM activity at online version, at doi:https://doi.org/10.1016/j.apcatb.2020.118859.
lower reaction temperature condition but also higher deactivation
possibility through oxidation. The possibility of carbon deposition over References
larger particles of Co, which are dominated by terrace sites, is strongly
dependent on the reaction temperature, and higher temperatures will [1] T. Yabe, Y. Sekine, Methane conversion using carbon dioxide as an oxidizing agent:
inhibit the coke formation. Based on these understandings, we further a review, Fuel Process. Technol. 181 (2018) 187–198.
[2] S.R. Foit, I.C. Vinke, L.G.J. de Haart, R.-A. Eichel, Power-to-Syngas – eine
propose below some suggestions for the modification of Co-based cat- Schlüsseltechnologie für die Umstellung des Energiesystems? Angew. Chem. 129
alysts. (2017) 5488–5498.
In order to obtain balanced activity and stability of the DRM cata- [3] S. Golestan, A.A. Mirzaei, H. Atashi, Fischer-Tropsch synthesis over an iron-cobalt-
manganese (ternary) nanocatalyst prepared by hydrothermal procedure: effects of
lysts, we recently found that a good dopant should give rise to a proper
nanocatalyst composition and operational conditions, Int. J. Hydrog. Energy 42
adsorption energy of carbon to ensure that the C* formation process, (2017) 9816–9830.
e.g. CH4 dissociation, is rate controlling to improve the carbon re- [4] P.F. Cao, S. Adegbite, H.T. Zhao, E. Lester, T. Wu, Tuning dry reforming of methane
for F-T syntheses: a thermodynamic approach, Appl. Energy 227 (2018) 190–197.
sistance, and relatively low dissociation barriers of CH4 and CO2 can be
[5] D.H. Mei, V.A. Glezakou, V. Lebarbier, L. Kovarik, H.Y. Wan, K.O. Albrecht,
achieved to maintain a good activity [75]. By doping or forming an M. Gerber, R. Rousseau, R.A. Dagle, Highly active and stable MgAl2O4-supported
alloy, the adsorption of O* at the Co step is regulated and the oxidation Rh and Ir catalysts for methane steam reforming: a combined experimental and
of the Co catalyst can be inhibited. For Co with large particles and more theoretical study, J. Catal. 316 (2014) 11–23.
[6] Y. Khani, Z. Shariatinia, F. Bahadoran, High catalytic activity and stability of
(111) facets, our simulation suggests that high reaction temperature ZnLaAlO4 supported Ni, Pt and Ru nanocatalysts applied in the dry, steam and
should be achieved to reduce carbon deposition. combined dry-steam reforming of methane, Chem. Eng. J. 299 (2016) 353–366.
[7] N. Kopfle, T. Gotsch, M. Grunbacher, E.A. Carbonio, M. Havecker, A. Knop-Gericke,
L. Schlicker, A. Doran, D. Kober, A. Gurlo, S. Penner, B. Klotzer, Zirconium-assisted
4. Conclusions activation of palladium to boost syngas production by methane dry reforming,
Angew. Chem.-Int. Edit. 57 (2018) 14613–14618.
In this work, the reactivity and reaction pathway of DRM on Co [8] B. Abdullah, N.A.A. Ghani, D.V.N. Vo, Recent advances in dry reforming of methane
over Ni-based catalysts, J. Clean. Prod. 162 (2017) 170–185.
(111) and Co(211) were studied with a combined DFT calculation and [9] A.E. Castro Luna, M.E. Iriarte, Carbon dioxide reforming of methane over a metal
microkinetic modeling approach. We found that the step site over Co modified Ni-Al2O3 catalyst, Appl. Catal. A Gen. 343 (2008) 10–15.
(211) is the active site for the DRM reaction. On Co(111), the C + O [10] S. Damyanova, B. Pawelec, K. Arishtirova, J.L.G. Fierro, Ni-based catalysts for re-
forming of methane with CO2, Int. J. Hydrog. Energy 37 (2012) 15966–15975.
and CH + O paths are the preferred reaction paths, while the C + O

7
S. Chen, et al. Applied Catalysis B: Environmental 270 (2020) 118859

[11] Y. Lou, M. Steib, Q. Zhang, K. Tiefenbacher, A. Horvath, A. Jentys, Y. Liu, dependent electrocatalytic reduction of CO2 over Pd nanoparticles, J. Am. Chem.
J.A. Lercher, Design of stable Ni/ZrO2 catalysts for dry reforming of methane, J. Soc. 137 (2015) 4288–4291.
Catal. 356 (2017) 147–156. [43] P. Zhao, Z. Cao, X. Liu, P. Ren, D.-B. Cao, H. Xiang, H. Jiao, Y. Yang, Y.-W. Li, X.-
[12] Z. Zuo, S. Liu, Z. Wang, C. Liu, W. Huang, J. Huang, P. Liu, Dry reforming of me- D. Wen, Morphology and reactivity evolution of HCP and FCC Ru nanoparticles
thane on single-site Ni/MgO catalysts: importance of site confinement, ACS Catal. under CO atmosphere, ACS Catal. 9 (2019) 2768–2776.
(2018) 9821–9835. [44] G. Kresse, J. Hafner, Ab-initio molecular-dynamics simulation of the liquid-metal
[13] S. Wen, M. Liang, J. Zou, S. Wang, X. Zhu, L. Liu, Z.-j. Wang, Synthesis of a SiO2 amorphous-semiconductor transition in germanium, Phys. Rev. B 49 (1994)
nanofibre confined Ni catalyst by electrospinning for the CO2 reforming of methane, 14251–14269.
J. Mater. Chem. A 3 (2015) 13299–13307. [45] G. Kresse, J. Furthmuller, Efficient iterative schemes for ab initio total-energy cal-
[14] S. Kawi, Y. Kathiraser, J. Ni, U. Oemar, Z.W. Li, E.T. Saw, Progress in synthesis of culations using a plane-wave basis set, Phys. Rev. B 54 (1996) 11169–11186.
highly active and stable nickel-based catalysts for carbon dioxide reforming of [46] G. Kresse, J. Furthmuller, Efficiency of ab-initio total energy calculations for metals
methane, ChemSusChem 8 (2015) 3556–3575. and semiconductors using a plane-wave basis set, Comput. Mater. Sci. 6 (1996)
[15] C.J. Liu, J.Y. Ye, J.J. Jiang, Y.X. Pan, Progresses in the Preparation of Coke Resistant 15–50.
Ni-based Catalyst for Steam and CO2 Reforming of Methane, ChemCatChem 3 [47] P.E. Blochl, Projector augmented-wave method, Phys. Rev. B 50 (1994)
(2011) 529–541. 17953–17979.
[16] J.Y. Lu, Y. Guo, Q.R. Liu, G.Z. Han, Z.J. Wang, Co-based catalysts for carbon dioxide [48] J. Wellendorff, K.T. Lundgaard, A. Mogelhoj, V. Petzold, D.D. Landis, J.K. Norskov,
reforming of methane to synthesis gas, Prog. Chem. 29 (2017) 1471–1479. T. Bligaard, K.W. Jacobsen, Density functionals for surface science: exchange-cor-
[17] J.X. Liu, H.Y. Su, D.P. Sun, B.Y. Zhang, W.X. Li, Crystallographic dependence of CO relation model development with Bayesian error estimation, Phys. Rev. B 85
activation on cobalt catalysts: HCP versus FCC, J. Am. Chem. Soc. 135 (2013) (2012) 23.
16284–16287. [49] P.P. Wu, B. Yang, Theoretical insights into the promotion effect of subsurface boron
[18] A.W. Budiman, S.H. Song, T.S. Chang, C.H. Shin, M.J. Choi, Dry reforming of me- for the selective hydrogenation of CO to methanol over Pd catalysts, Phys. Chem.
thane over cobalt catalysts: a literature review of catalyst development, Catal. Surv. Chem. Phys. 18 (2016) 21720–21729.
Asia 16 (2012) 183–197. [50] K.L. Wang, B. Yang, Theoretical understanding on the selectivity of acrolein hy-
[19] K. Nagaoka, K. Takanabe, K. Aika, Modification of Co/ TiO2 for dry reforming of drogenation over silver surfaces: the non-Horiuti-Polanyi mechanism is the key,
methane at 2 MPa by Pt, Ru or Ni, Appl. Catal. A-Gen. 268 (2004) 151–158. Catal. Sci. Technol. 7 (2017) 4024–4033.
[20] K. Omata, N. Nukui, T. Hottai, M. Yamada, Cobalt-magnesia catalyst by oxalate co- [51] P.P. Wu, B. Yang, Significance of Surface Formate Coverage on the Reaction
precipitation method for dry reforming of methane under pressure, Catal. Commun. Kinetics of Methanol Synthesis from CO2 Hydrogenation over Cu, ACS Catal. 7
5 (2004) 771–775. (2017) 7187–7195.
[21] D.S. Jose-Alonso, M.J. Illan-Gomez, M.C. Roman-Martinez, Low metal content Co [52] K.R. Yang, B. Yang, Identification of the active and selective sites over a single Pt
and Ni alumina supported catalysts for the CO2 reforming of methane, Int. J. atom-alloyed Cu catalyst for the hydrogenation of 1,3-Butadiene: a combined DFT
Hydrog. Energy 38 (2013) 2230–2239. and microkinetic modeling study, J. Phys. Chem. C 122 (2018) 10883–10891.
[22] F. Mirzaei, M. Rezaei, F. Meshkani, Z. Fattah, Carbon dioxide reforming of methane [53] P.P. Wu, J. Zaffran, B. Yang, Role of surface species interactions in identifying the
for syngas production over Co-MgO mixed oxide nanocatalysts, J. Ind. Eng. Chem. reaction mechanism of methanol synthesis from CO2 hydrogenation over inter-
21 (2015) 662–667. metallic Pdln(310) Steps, J. Phys. Chem. C 123 (2019) 13615–13623.
[23] J. Cheng, P. Hu, P. Ellis, S. French, G. Kelly, C.M. Lok, Some understanding of [54] D.Y. Xu, P.P. Wu, B. Yang, Essential role of water in the autocatalysis behavior of
Fischer–Tropsch synthesis from density functional theory calculations, Top. Catal. methanol synthesis from CO2 hydrogenation on Cu: a combined DFT and micro-
53 (2010) 326–337. kinetic modeling study, J. Phys. Chem. C 123 (2019) 8959–8966.
[24] R.A. van Santen, M.M. Ghouri, S. Shetty, E.M.H. Hensen, Structure sensitivity of the [55] H. Liu, J. Liu, B. Yang, Modeling the effect of surface CO coverage on the electro-
Fischer–Tropsch reaction; molecular kinetics simulations, Catal. Sci. Technol. 1 catalytic reduction of CO2 to CO on Pd surfaces, Phys. Chem. Chem. Phys. 21 (2019)
(2011) 891–911. 9876–9882.
[25] C.Q. Chen, Y.Y. Zhan, J.K. Zhou, D.L. Li, Y.J. Zhang, X.Y. Lin, L.L. Jiang, Q. Zheng, [56] A. Alavi, P.J. Hu, T. Deutsch, P.L. Silvestrelli, J. Hutter, CO oxidation on Pt(111): an
Cu/CeO2 catalyst for water-gas shift reaction: effect of CeO2 pretreatment, ab initio density functional theory study, Phys. Rev. Lett. 80 (1998) 3650–3653.
ChemPhysChem 19 (2018) 1448–1455. [57] Z.P. Liu, P. Hu, General rules for predicting where a catalytic reaction should occur
[26] H.Y. Wang, E. Ruckenstein, CO2 reforming of CH4 over Co/MgO solid solution on metal surfaces: a density functional theory study of C-H and C-O bond breaking/
catalysts - effect of calcination temperature and Co loading, Appl. Catal. A-Gen. 209 making on flat, stepped, and kinked metal surfaces, J. Am. Chem. Soc. 125 (2003)
(2001) 207–215. 1958–1967.
[27] E. Ruckenstein, H.Y. Wang, Carbon deposition and catalytic deactivation during [58] A. Michaelides, Z.P. Liu, C.J. Zhang, A. Alavi, D.A. King, P. Hu, Identification of
CO2 reforming of CH4 over Co/gamma-Al2O3 catalysts, J. Catal. 205 (2002) general linear relationships between activation energies and enthalpy changes for
289–293. dissociation reactions at surfaces, J. Am. Chem. Soc. 125 (2003) 3704–3705.
[28] D. San-Jose-Alonso, J. Juan-Juan, M.J. Illan-Gomez, M.C. Roman-Martinez, Ni, Co [59] H. Schäfer, J. Donohue, The Structures of the Elements, John Wiley & Sons, New
and bimetallic Ni-Co catalysts for the dry reforming of methane, Appl. Catal. A-Gen. York, Sydney, Toronto, 1974 436 S., Preis: £ 12.—, 79 (1975) 110–110.
371 (2009) 54–59. [60] X. Liu, L. Sun, W.-Q. Deng, Theoretical investigation of CO2 adsorption and dis-
[29] M.F.F. Sukri, M. Khavarian, A.R. Mohamed, Effect of cobalt loading on suppression sociation on low index surfaces of transition metals, J. Phys. Chem. C 122 (2018)
of carbon formation in carbon dioxide reforming of methane over Co/MgO catalyst, 8306–8314.
Res. Chem. Intermed. 44 (2018) 2585–2605. [61] M. Methfessel, A.T. Paxton, High-precision sampling for Brillouin-zone integration
[30] E. Ruckenstein, H.Y. Wang, Carbon dioxide reforming of methane to synthesis gas in metals, Phys. Rev. B 40 (1989) 3616–3621.
over supported cobalt catalysts, Appl. Catal. A-Gen. 204 (2000) 257–263. [62] A.A. Peterson, F. Abild-Pedersen, F. Studt, J. Rossmeisl, J.K. Norskov, How copper
[31] Y. Guo, J.Y. Lu, Q.R. Liu, X.L. Bai, L.J. Gao, W.X. Tu, Z.J. Wang, Carbon dioxide catalyzes the electroreduction of carbon dioxide into hydrocarbon fuels, Energy
reforming of methane over cobalt catalysts supported on hydrotalcite and metal Environ. Sci. 3 (2010) 1311–1315.
oxides, Catal. Commun. 116 (2018) 81–84. [63] R. Christensen, H.A. Hansen, T. Vegge, Identifying systematic DFT errors in cata-
[32] C.A. Singh, L. Bansal, P. Tiwari, V.V. Krishnan, Strong metal support interactions of lytic reactions, Catal. Sci. Technol. 5 (2015) 4946–4949.
infiltrated Ni with TiO2 in a porous YSZ anode matrix - a possible method for Ni- [64] A.J. Medford, C. Shi, M.J. Hoffmann, A.C. Lausche, S.R. Fitzgibbon, T. Bligaard,
stabilization, in: S.C. Singhal, H. Yokokawa (Eds.), Solid Oxide Fuel Cells 11, J.K. Norskov, CatMAP: a software package for descriptor-based microkinetic map-
Electrochemical Soc Inc, Pennington, 2009, pp. 1897–1904. ping of catalytic trends, Catal. Lett. 145 (2015) 794–807.
[33] L. Yu, Y. Liu, F. Yang, J. Evans, J.A. Rodriguez, P. Liu, CO oxidation on gold-sup- [65] P. Wu, B. Yang, Intermetallic PdIn catalyst for CO2 hydrogenation to methanol:
ported Iron oxides: new insights into strong oxide–metal interactions, J. Phys. mechanistic studies with a combined DFT and microkinetic modeling method,
Chem. C 119 (2015) 16614–16622. Catal. Sci. Technol. 9 (2019) 6102–6113.
[34] J.-H. Park, S. Yeo, T.-S. Chang, Effect of supports on the performance of Co-based [66] P. Wang, B. Yang, Influence of surface strain on activity and selectivity of Pd-based
catalysts in methane dry reforming, J. CO2 UTIL. 26 (2018) 465–475. catalysts for the hydrogenation of acetylene: a DFT study, Chin. J. Catal. 39 (2018)
[35] T. Nishizawa, K. Ishida, The Co (Cobalt) System, Bull. Alloy. Phase Diagr. 4 (1983) 1493–1499.
387. [67] C. Stegelmann, A. Andreasen, C.T. Campbell, Degree of rate control: how much the
[36] J. Gong, L.L. Wang, Y. Liu, J.H. Yang, Z.G. Zong, Structural and magnetic properties energies of intermediates and transition states control rates, J. Am. Chem. Soc. 131
of hcp and fcc Ni nanoparticles, J. Alloys. Compd. 457 (2008) 6–9. (2009) 8077–8082.
[37] O. Kitakami, H. Sato, Y. Shimada, F. Sato, M. Tanaka, Size effect on the crystal [68] C.T. Campbell, The degree of rate control: a powerful tool for catalysis research,
phase of cobalt fine particles, Phys. Rev. B 56 (1997) 13849–13854. ACS Catal. 7 (2017) 2770–2779.
[38] R. Lizarraga, F. Pan, L. Bergqvist, E. Holmstrom, Z. Gercsi, L. Vitos, First principles [69] C.A. Wolcott, A.J. Medford, F. Studt, C.T. Campbell, Degree of rate control ap-
theory of the hcp-fcc phase transition in cobalt, Sci. Rep. 7 (2017) 8. proach to computational catalyst screening, J. Catal. 330 (2015) 197–207.
[39] K. Takanabe, K. Nagaoka, K. Nariai, K. Aika, Influence of reduction temperature on [70] S. Chen, B. Yang, Theoretical understandings on the unusual selectivity of 1,3-
the catalytic behavior of Co/TiO2 catalysts for CH4/ CO2 reforming and its relation Butadiene hydrogenation to butenes over gold catalysts, Catal. Today (2018).
with titania bulk crystal structure, J. Catal. 230 (2005) 75–85. [71] B. Wang, S. Chen, J. Zhang, S. Li, B. Yang, Propagating DFT uncertainty to me-
[40] H. Ay, D. Uner, Dry reforming of methane over CeO2 supported Ni, Co and Ni-Co chanism determination, degree of rate control, and coverage analysis: the kinetics
catalysts, Appl. Catal. B-Environ. 179 (2015) 128–138. of dry reforming of methane, J. Phys. Chem. C 123 (2019) 30389–30397.
[41] Y.J. Wong, M.K. Koh, M. Khavarian, A.R. Mohamed, Investigation on cobalt alu- [72] Y.A. Zhu, D. Chen, X.G. Zhou, W.K. Yuan, DFT studies of dry reforming of methane
minate as an oxygen carrier catalyst for dry reforming of methane, Int. J. Hydrog. on Ni catalyst, Catal. Today 148 (2009) 260–267.
Energy 42 (2017) 28363–28376. [73] L. Foppa, M.C. Silaghi, K. Larmier, A. Comas-Vives, Intrinsic reactivity of Ni, Pd and
[42] D. Gao, H. Zhou, J. Wang, S. Miao, F. Yang, G. Wang, J. Wang, X. Bao, Size- Pt surfaces in dry reforming and competitive reactions: insights from first principles

8
S. Chen, et al. Applied Catalysis B: Environmental 270 (2020) 118859

calculations and microkinetic modeling simulations, J. Catal. 343 (2016) 196–207. Petrochem. Res. 4 (2014) 137–144.
[74] C. Fan, Y.A. Zhu, M.L. Yang, Z.J. Sui, X.G. Zhou, D. Chen, Density functional theory- [80] J.P. Tessonnier, D.S. Su, Recent progress on the growth mechanism of carbon na-
assisted microkinetic analysis of methane dry reforming on Ni catalyst, Ind. Eng. notubes: a review, ChemSusChem 4 (2011) 824–847.
Chem. Res. 54 (2015) 5901–5913. [81] P. Sautet, F. Cinquini, Surface of metallic catalysts under a pressure of hydrocarbon
[75] S. Chen, J. Zaffran, B. Yang, Descriptor design in the computational screening of Ni- molecules: metal or carbide? ChemCatChem 2 (2010) 636–639.
based catalysts with balanced activity and stability for the dry reforming of me- [82] K. Nagaoka, K. Takanabe, K. Aika, Influence of the reduction temperature on cat-
thane reaction, ACS Catal. 10 (2020) 3074–3083. alytic activity of Co/TiO2 (anatase-type) for high pressure dry reforming of me-
[76] B. AlSabban, L. Falivene, S.M. Kozlov, A. Aguilar-Tapia, S. Ould-Chikh, J.- thane, Appl. Catal. A-Gen. 255 (2003) 13–21.
L. Hazemann, L. Cavallo, J.-M. Basset, K. Takanabe, In-operando elucidation of [83] J.L. Ewbank, L. Kovarik, C.C. Kenvin, C. Sievers, Effect of preparation methods on
bimetallic CoNi nanoparticles during high-temperature CH4/ CO2 reaction, Appl. the performance of Co/Al2O3 catalysts for dry reforming of methane, Green Chem.
Catal. B 213 (2017) 177–189. 16 (2014) 885–896.
[77] J. Horlyck, C. Lawrey, E.C. Lovell, R. Amal, J. Scott, Elucidating the impact of Ni [84] L. Ji, S. Tang, H.C. Zeng, J. Lin, K.L. Tan, CO2 reforming of methane to synthesis gas
and Co loading on the selectivity of bimetallic NiCo catalysts for dry reforming of over sol–gel-made Co/γ-Al2O3 catalysts from organometallic precursors, Appl.
methane, Chem. Eng. J. 352 (2018) 572–580. Catal. A Gen. 207 (2001) 247–255.
[78] M.S. Aw, I.G. Osojnik Črnivec, P. Djinović, A. Pintar, Strategies to enhance dry [85] D. San José-Alonso, M.J. Illán-Gómez, M.C. Román-Martínez, K and Sr promoted Co
reforming of methane: synthesis of ceria-zirconia/nickel–cobalt catalysts by freeze- alumina supported catalysts for the CO2 reforming of methane, Catal. Today 176
drying and NO calcination, Int. J. Hydrog. Energy 39 (2014) 12636–12647. (2011) 187–190.
[79] X. Du, L.J. France, V.L. Kuznetsov, T. Xiao, P.P. Edwards, H. AlMegren, A. Bagabas, [86] Z. Hou, T. Yashima, Supported Co catalysts for methane reforming with CO2, React.
Dry reforming of methane over ZrO2-supported Co–Mo carbide catalyst, Appl. Kinet. Catal. Lett. 81 (2004) 153–159.

You might also like