Effect of Rutile Content on the Catalytic Performance of Ru_TiO2 Catalyst for Low-Temperature CO2 Methanation

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

pubs.acs.

org/journal/ascecg Research Article

Effect of Rutile Content on the Catalytic Performance of Ru/TiO2


Catalyst for Low-Temperature CO2 Methanation
Zhiying Zhao, Qiaorong Jiang, Qiuxiang Wang, Mingzhi Wang, Jiachang Zuo, Hanming Chen,
Qin Kuang,* and Zhaoxiong Xie*
Cite This: ACS Sustainable Chem. Eng. 2021, 9, 14288−14296 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via INDIAN INST OF TECH MADRAS on April 14, 2024 at 02:31:02 (UTC).

ABSTRACT: Hydrogenation of carbon dioxide (CO2) to methane (CH4) bears the


potential not only to alleviate excessive emission of CO2 but to produce clean energy for
direct use. However, due to the high temperature required for the activation of CO2, it is
necessary to develop a catalyst with high activity and high CH4 selectivity at low reaction
temperature for CO2 methanation. In this study, we synthesized titanium dioxide (TiO2)
supports with different rutile contents by calcination and investigated the relationship
between the rutile content of TiO2 supports in Ru/TiO2 catalysts and their performance
for low-temperature CO2 methanation. The characterization results indicated that Ru
species grew preferentially on the rutile phase of the support TiO2, and the interaction
between Ru species and TiO2 became stronger with increasing rutile content. In addition,
the concentration of oxygen vacancies on the support, which would provide more
adsorption sites for CO2, increased with the decreasing rutile content. The best reactivity
was achieved with a rutile content between 15% and 30%. This work included a thorough
investigation on the effect of rutile content on the performance of low-temperature CO2 methanation and provided guidance for the
preparation of Ru/TiO2 catalyst with high activity.
KEYWORDS: Ru/TiO2, CO2 methanation, Rutile content, Interaction, Oxygen vacancy

■ INTRODUCTION
The catalytic hydrogenation of carbon dioxide (CO2) to
deactivation by carbon deposition than Ni and less expensive
than Rh.11−14
Previous studies suggested that the metal sites in supported
methane (CH4) was first discovered by Sabatier in 1902 and is
metal/oxide catalysts were able to dissociate hydrogen (H2),
termed CO2 methanation or the Sabatier reaction.1 Since the while the oxide and the metal−oxide interface were responsible
industrial revolution, the overuse of fossil fuels has exacerbated for the adsorption and activation of CO2.15−17 Therefore, the
environmental and energy-related problems. To reduce the type of oxide support and the construction of the metal−oxide
adverse impact of excessive CO2 emission, CO2 methanation interface were also important to promote the performance of
has received renewed attention in recent decades.2−4 Chemical CO2 methanation.18 Titanium dioxide (TiO2) has been widely
conversion of CO2 to CH4 could not only reduce CO2 used as the catalyst support in thermocatalysis due to its low
emission but also generate clean energy that can be used price, high stability, and abundant oxygen vacancies.19−21 Liu
directly at the same time.5 Therefore, CO2 methanation was and co-workers22 highlighted the importance of oxygen
considered to be one of the most promising energy storage vacancies for the adsorption of CO2, showing that CO2
strategies. For an efficient activation of the C−O bond in CO2, bonds weakly on the surface of a silica-supported platinum
reaction temperatures above 350 °C are usually reported in catalyst (Pt/SiO2) due to the nonreducibility of SiO2. In
most of the literature for CO2 methanation.6,7 However, the contrast, Pt/TiO2 with more oxygen vacancies exhibited higher
high reaction temperature requires an actual increase in energy activity for the conversion of CO2 to CO. In the recent 20
consumption and commonly causes serious aggregation of years, Ru/TiO2 has been reported as a typical catalyst for CO2
active metal species on catalysts.8 Therefore, designing a methanation, while the phase of TiO2 was identified as one of
catalyst with high activity and high CH4 selectivity at low
reaction temperatures rationally is still one of the major goals Received: August 15, 2021
in CO2 methanation.7 Up to now, supported metal-based Revised: October 5, 2021
catalysts (e.g., Ni, Ru, and Rh) have been reported as potential Published: October 13, 2021
catalysts for this reaction in a large number of studies.9,10
Among them, Ru-based catalysts are more promising for CO2
methanation applications since they are less conducive to

© 2021 American Chemical Society https://doi.org/10.1021/acssuschemeng.1c05565


14288 ACS Sustainable Chem. Eng. 2021, 9, 14288−14296
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

the significant factors determining the catalytic performance in High-angle annular dark-field scanning transmission electron
CO2 hydrogenation.23−26 In earlier research, the existence of a microscopy (HAADF-STEM), high-resolution transmission electron
strong interaction between Ru species and the rutile phase of microscopy (HRTEM), and element mapping analyses were
performed with a Phillips FEI Tecnai 20 microscope operated at
TiO2 was discovered, which could prohibit the aggregation of
200 kV; selected area electron diffraction (SAED) was performed on a
Ru species to a great extent.27,28 However, the existing reports JEOL JEM-2100 microscope. The morphology and size of samples
only focused on the advantages of rutile TiO2 but did not were characterized by a Hitachi S-4800 scanning electron microscope
conduct detailed studies on the effect of introducing the (SEM).
anatase phase of TiO2 to the catalyst on CO2 methanation CO2 (and CO) temperature-programmed desorption (CO2-TPD
performance at low reaction temperature. As we all know, and CO-TPD) experiments were performed on self-built equipment
rutile TiO2 provides fewer oxygen vacancies due to its (Hiden Analytical QIC-20 quadrupole mass spectrometer). In a
comparatively lower reducibility and smaller specific surface typical experiment process of CO2-TPD, an amount of 60 mg of
area compared to the anatase TiO2 phase.29,30 However, since catalyst was reduced at 400 °C for 1 h under an atmosphere of 5%
H2/Ar and then cooled to room temperature in He (99.999%)
oxygen vacancies significantly contribute to CO2 hydro- atmosphere. Afterward, a stream of CO2 (99.99%) was introduced
genation, the influence of the phase content on low- into the reactor for 20 min. After the adsorption of CO2, the catalyst
temperature CO2 methanation requires further exploration. was scrubbed with He (99.999%) for 30 min to remove physically
In this work, we prepared a series of Ru/TiO2 catalysts with adsorbed CO2. Finally, the catalyst was heated in He (99.999%) to
different phase contents through a simple impregnation 500 °C. Online mass spectrometry was employed to track ion
method and high-temperature calcination. It was found that, fragments including CO2 (m/z = 44). In the experiment process of
at a rutile content in the range of 15−30%, both the interaction CO-TPD, the adsorption gas was changed to 1% CO/He, while
between Ru species and the TiO 2 support and the keeping other conditions unchanged. Online mass spectrometry was
employed to track ion fragments including CO (m/z = 28).
concentration of oxygen vacancies were moderate and suitable In situ diffuse reflectance infrared Fourier transform spectroscopy
for CO2 methanation at low reaction temperature. This study (DRIFTS) was conducted on a Thermo Nicolet 6700 spectrometer
provides a new perspective on the influence of the TiO2 phase equipped with ZnSe filters and a high-temperature reaction chamber.
on low-temperature CO2 methanation. Before the in situ CO2 methanation DRIFTS experiment, the catalyst


was reduced in a 5% H2/Ar flow at 400 °C for 1 h and then cooled to
EXPERIMENTAL SECTION 300 °C in a reducing atmosphere. After that, a stream of 15% CO2/
45% H2/N2 was introduced into the chamber and reacted for 5 min.
Materials and Chemicals. Ethanol and hydrogen peroxide Finally, the 15% CO2/45% H2/N2 flow was switched to 5% H2/Ar for
(H2O2) aqueous solution (30% w/v) were purchased from 5 min. The spectra (64 scans) were collected with a resolution of 4
Sinopharm Chemical Reagent Co., Ltd.; ruthenium(III) chloride cm−1.
(RuCl3) (99.9%) was purchased from J&K Scientific Ltd.; and The dispersion of Ru species (D) was calculated by eq 132
titanium isopropoxide [Ti(OC3H7)4] (99.9%) was purchased from
Energy Chemical Co., Ltd. All chemicals were used without further 6(v /a) 6Mns
D= =
purification. d ρNAd (1)
Preparation of TiO2 Supports. The TiO2 support was
synthesized by the reported sol−gel method.31 The preparation where v denotes the volume occupied by a Ru atom (Å3), a denotes
process comprised the following steps: (1) 3 mL of Ti(OC3H7)4 was the polycrystalline surface occupied by a Ru atom (Å2), d denotes the
dispersed in 10 mL of ethanol, and then 50 mL of distilled water was mean particle size (Å), M denotes the atomic mass (g mol−1), ns
added. After stirring for 20 min, the produced white precipitate was denotes the number of surface atoms per area unit (m−2), ρ denotes
centrifuged (9500 r/min) and washed with water three times to the density (g cm−3), and NA denotes Avogadro’s number (6.022 ×
remove residual ethanol. (2) The white product was slowly added to 1023 mol−1). In the case of a hexagonal closed-packed (hcp) Ru(001)
15 mL of aqueous H2O2 solution and stirred continuously until it surface, eq 1 was transformed to
dissolved to form an orange solution, which then changed into a 12.9
transparent orange sol. Finally, the sol slowly thickened to form a gel. D=
d (2)
(3) The gel was dried under an infrared lamp and then calcined at 500
°C (5 °C/min) for 3 h in air. The obtain TiO2 support was denoted Catalytic Tests. CO2 methanation reactions were performed in a
as TiO2-A. (4) The TiO2-A samples were calcined at 600, 700, or 800 fixed-bed reactor at atmospheric pressure. An amount of 50 mg of
°C for 3 h, and the obtained supports were denoted as TiO2-600, catalyst was loaded in a quartz tube. The reaction gas consisted of H2
TiO2-700, or TiO2-800, respectively. (27 mL/min), CO2 (9 mL/min), and N2 (24 mL/min). Before the
Preparation of Ru/TiO2 Catalysts. Ru/TiO2 catalysts were reaction test, the catalyst was reduced at 400 °C for 1 h in 20% H2/N2
prepared by an impregnation method. In a typical synthetic process of (60 mL/min). A specific reaction rate was determined under different
Ru/TiO2-A, 300 mg of TiO2-A was impregnated in 0.663 mL of conditions, using 10 mg of catalyst and a reaction gas composed of H2
aqueous RuCl3 solution (Ru: 9.0469 mg/mL) and then stirred (250 (45 mL/min), CO2 (15 mL/min), and N2 (40 mL/min) to control
r/min) at room temperature for 10 h. The sample was dried at 60 °C the reaction at a low CO2 conversion. CO2 conversion was calculated
for 12 h and then calcined at 500 °C (5 °C/min) for 2 h in air. by the following equation:
Similarly, Ru/TiO2-600, Ru/TiO2-700, and Ru/TiO2-800 catalysts COout + CH4 out
were prepared by replacing TiO2-A with TiO2-600, TiO2-700, and Conv CO2 = × 100%
TiO2-800, respectively. The theoretical loading of Ru was 2%. COout + CH4 out + CO2 out (3)
Characterization of Samples. Powder X-ray diffraction (XRD)
CO selectivity and CH4 selectivity were calculated as the following
was performed on a Rigaku Ultima IV X-ray diffractometer with Cu
equations
Kα (λ = 0.154 06 nm) radiation. The rutile content of each catalyst
was obtained by Rietveld refinement. X-ray photoelectron spectros- COout
copy (XPS) was performed on a PHI Quantum-2000 instrument. The Sel CO = × 100%
COout + CH4 out (4)
binding energies of all elements were calibrated against the C 1s peak
at 284.8 eV. The specific surface area (SBET) of the sample was CH4 out
measured by N2 adsorption−desorption isotherms at −196 °C using a Sel CH4 = × 100%
Micromeritics Tristar 3020. COout + CH4 out (5)

14289 https://doi.org/10.1021/acssuschemeng.1c05565
ACS Sustainable Chem. Eng. 2021, 9, 14288−14296
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

where CO2 out, COout, and CH4 out represent the molar amounts of indicated that only the anatase phase (ICDD PDF: 01-071-
CO2, CO, and CH4 in the effluent, respectively.


1167) was detected in TiO2-A, while a small amount of rutile
phase (ICDD PDF: 01-076-0318) was detected in TiO2-600
RESULTS AND DISCUSSION and its content gradually increased with the second calcination
Figure 1 shows the composition and morphology of TiO2-A temperature. The contents of rutile phase in different samples,
obtained via the sol−gel method and other TiO2 supports (i.e., determined by Rietveld refinement of XRD data, are listed in
Table S1 (SI). The content of rutile phase in TiO2-600 was
only 15−30%, while that in TiO2-800 exceeded 94%. It is
worth noting that the increasing trends of rutile content and
particle size with calcination temperature are basically the
same, as shown in Figure 1f. This indicates that those particles
with relatively larger sizes in the sample are rutile TiO2.
Ru species was deposited onto the mentioned four TiO2
supports through the impregnation method. The methanation
performance of the obtained Ru/TiO2 catalysts determined at
200, 250, and 300 °C are shown in Figure 2. The CO2
conversions over these catalysts increased with the reaction
temperature (Figure 2a). At all reaction temperatures, the CO2
conversion first increased and then decreased with the increase
in rutile content of the TiO2 support, and the highest
conversion was achieved on the catalyst Ru/TiO2-600 with a
rutile content of 15−30%. For the same catalyst, the CO2
conversion reached 21% at 300 °C and remained stable for 80
h (Figure S2, SI). Interestingly, no significant difference in the
CH4 selectivity was observed for the different catalysts (Figure
2b), which exceeded 99% at all investigated reaction
temperatures. The catalysts were further characterized to
reveal the influence of rutile content on CO2 conversion.
The dispersion status of Ru species on Ru/TiO2 catalysts
obtained by calcination at 500 °C in air were determined by
TEM. As illustrated in Figure 3, some nanoparticles with dark
contrast were observed on the Ru/TiO2 catalysts, which were
Figure 1. (a−d) SEM images of samples, (e) XRD patterns of
attributed to RuOx species. Notably, RuOx nanoparticles on
samples, and (f) particle size and rutile content of samples. the Ru/TiO2-A, Ru/TiO2-600, and Ru/TiO2-700 catalysts are
of similar size, which are 2.5 ± 0.7, 2.5 ± 0.4 and 2.2 ± 0.4 nm,
respectively. By contrast, the RuOx particle size was slightly
TiO2-600, TiO2-700, and TiO2-800) obtained by subsequent larger on Ru/TiO2-800 (3.0 ± 0.8 nm). Due to their small
calcination treatment of TiO2-A at different temperatures.31 sizes and low loading, diffraction peaks related to Ru species
The size of TiO2 particles was about 20−30 nm in TiO2-A were not detected in the XRD patterns (Figure S3, SI). To
(Figure 1a). With the increase of the second calcination explore the reasons why CO2 conversion over Ru/TiO2-A, Ru/
temperature, the size of TiO2 particles increased gradually to TiO2-600, and Ru/TiO2-700 catalysts first increased and then
30−50 nm for TiO2-600, 50−70 nm for TiO2-700, and 100− decreased, these three catalysts were selected for in-depth
150 nm for TiO2-800 (Figure 1b−d). According to the study.
measured N2 adsorption−desorption isotherms [Figure S1, Figure 4 shows TEM characterization results of Ru/TiO2
Supporting Information (SI)], specific surface areas of TiO2-A, catalysts after reduction at 400 °C. Ru species were still highly
TiO2-600, TiO2-700, and TiO2-800 were 38, 15, 13, and 2 m2/ dispersed on TiO2 particles in Ru/TiO2-A with a pure anatase
g, respectively (Table S1, SI). XRD patterns (Figure 1e) phase, but the sizes were significantly increased (Figure 4a).

Figure 2. CO2 methanation on Ru/TiO2 catalysts (2% Ru) at different temperatures; m = 50 mg, reduction temperature = 400 °C, feed: 15% CO2
+ 45% H2 in N2 (total flow rate = 60 mL/min). (a) CO2 conversion and (b) CH4 selectivity.

14290 https://doi.org/10.1021/acssuschemeng.1c05565
ACS Sustainable Chem. Eng. 2021, 9, 14288−14296
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 3. TEM images and corresponding particle size distributions of catalysts obtained by calcination at 500 °C in air: (a) Ru/TiO2-A, (b) Ru/
TiO2-600, (c) Ru/TiO2-700, and (d) Ru/TiO2-800.

Figure 4. TEM images of (a) Ru/TiO2-A, (b) Ru/TiO2-600, and (c) Ru/TiO2-700 catalysts after reduction at 400 °C in H2/N2. (d) HAADF-
STEM images of Ru/TiO2-600 and the corresponding elemental mapping images of Ru/TiO2-600 catalyst after reduction at 400 °C. Ru
nanoparticles are indicated by white arrows.

Interestingly, Ru species were found to grow preferentially on spacing (0.32 nm) of rutile (110). According to the dark-field
the TiO2 particles of larger size (dark lines indicated by white TEM image corresponding to these rutile (110) diffraction
arrows in Figure 4b) in Ru/TiO2-600 after reduction in H2. A spots (Figure 5e), Ru species (indicated by white arrows in
similar phenomenon was observed on Ru/TiO2-700, which Figure 5c) indeed grew preferentially on the rutile particles.
was also composed of mixed phases of rutile and anatase The average sizes of Ru species (Figure S5, SI) on Ru/TiO2-A,
[Figures 4c and S4 (SI)]. This selective enrichment behavior Ru/TiO2-600, and Ru/TiO2-700 after reduction at 400 °C in
of Ru species was seen more clearly in the HAADF-STEM H2 were 4.5, 3.4, and 3.0 nm, and the corresponding
image and corresponding elemental mapping images of Ru/ dispersions of Ru species calculated by eq 2 were 28.7%,
TiO2-600 (Figure 4d). 37.9%, and 43%, respectively. The observed increase in the
To explore the exact dispersion of Ru species on TiO2 dispersion of Ru species with increasing rutile content might
particles of different phases, HRTEM and SAED character-
indicate that rutile TiO2 could stabilize Ru species to some
izations were carried out on Ru/TiO2-600 using Ru/TiO2-A as
a reference. As visible in the HRTEM image of Ru/TiO2-A extent and then prevent them from aggregating. Also, it had
(Figure 5a), the Ru species were located on anatase particles been documented that RuO2 sinters heavily on anatase but
[TiO2-A(101), d = 0.35 nm]. In contrast, Ru species spreads and forms epitaxial layers on rutile.25,28 According to
preferentially grew on rutile particles [TiO2-R(110), d = 0.32 the literature and the above TEM characterization results, one
nm] in Ru/TiO2-600 (Figure 5b). Furthermore, Ru/TiO2-600 can infer that the Ru species on the Ru/TiO2 catalysts
was analyzed by SAED and HAADF-STEM techniques. As prepared in this work grew preferentially on rutile particles.
shown in Figure 5c,d, the value of 1/d for the diffraction spots This phenomenon was still observed on the catalysts after CO2
(indicated by the yellow arrow) in the SAED pattern recorded methanation reaction (Figure S6, SI). Meanwhile, the average
from a large-sized TiO2 particle with deposited Ru species was sizes of Ru species on Ru/TiO2-A, Ru/TiO2-600, and Ru/
3.077 1/nm, which was consistent with the lattice fringe TiO2-700 after CO2 methanation reaction were 4.3, 3.5, and
14291 https://doi.org/10.1021/acssuschemeng.1c05565
ACS Sustainable Chem. Eng. 2021, 9, 14288−14296
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 5. HRTEM images of (a) Ru/TiO2-A and (b) Ru/TiO2-600 catalysts reduced at 400 °C. The insets are the magnified pictures of marked
lattice fringes. (c) TEM image of Ru/TiO2-600 catalyst after reduction at 400 °C, (d) SAED pattern, and (e) dark-field TEM image of the catalyst
corresponding to the selected diffraction spots indicated by the yellow arrow in part d. Ru particles are indicated by white arrows.

Figure 6. C 1s and Ru 3d XPS spectra of (a) Ru/TiO2-A, (b) Ru/TiO2-600, and (c) Ru/TiO2-700 reduced at 400 °C in H2.

3.4 nm, respectively. This meant that the size of the catalyst TiO2-700 were 25% (Figure 6b) and 30% (Figure 6c),
did not change much after reaction. respectively. The results revealed that after high-temperature
XPS was employed to analyze the chemical states of Ru reduction, the predominant state of Ru species loaded on pure
species in Ru/TiO2 catalysts of different rutile content. The anatase was Ru0. When the rutile phase was present, the
binding energy of Ru 3d5/2 and the valence state of Ru species catalyst could still retain some oxidized Ru species, and the
are listed in Table S2 (SI). After calcination at 500 °C in air proportion of residual oxidized Ru species increased with the
(Figure S7, SI), only one peak at 280.2 eV (Ru 3d5/2) rutile content. As a possible explanation, the preferential
corresponding to Ruδ+ (0 < δ < 4) was observed on Ru/TiO2- growth of Ru species on rutile particles led to a stronger
A. By contrast, an additional peak (Ru 3d5/2) at higher binding interaction between the Ru and support and the observed
energy (281.8−282.0 eV) was present on Ru/TiO2-600 and higher average oxidation states of Ru species. Lin et al.27 also
Ru/TiO2-700, which was attributed to Ru4+. The proportion of reported that the Ru species on pure anatase support could be
Ru4+ species in the two catalysts were 16% and 18%, reduced to the metallic state at 400 °C, while the Ru species on
respectively.33 After reduction at 400 °C in H2 (Figure 6a) the pure rutile support could maintain their oxidation state due
or the CO2 methanation reaction (Figure S8a, SI), Ru/TiO2-A to the strong interaction between RuO2 and rutile.
only exhibited one Ru 3d5/2 peak assigned to Ru0 species According to the TEM and Ru-XPS characterization results,
(279.6 eV), which indicated that all Ru species in Ru/TiO2-A the increase of rutile content indeed led to a strong interaction
changed to the metallic state.34 However, in Ru/TiO2-600 and between TiO2 and Ru species, which partly prevented sintering
Ru/TiO2-700, a large number of Ru0 species (279.6−279.7 of the active sites and resulted in a higher reactivity of the
eV) was generated, but a considerable portion of oxidized Ru catalyst. However, while the rutile content of Ru/TiO2-700
species (281.0−282.0 eV) remained. After reduction at 400 °C was higher than that of Ru/TiO2-600, the reactivity of Ru/
in H2, the proportion of Ruδ+ species in Ru/TiO2-600 and Ru/ TiO2-700 was lower than that of Ru/TiO2-600. To explore the
14292 https://doi.org/10.1021/acssuschemeng.1c05565
ACS Sustainable Chem. Eng. 2021, 9, 14288−14296
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

Figure 7. (a) O 1s XPS spectra and (b) CO2-TPD curves of catalysts reduced at 400 °C.

Figure 8. In situ DRIFTS of the CO2 methanation reaction at 300 °C on Ru/TiO2-600 catalyst. (a) The catalyst was exposed to 15%CO2/45%H2/
N2 for 5 min and (c) the catalyst was scrubbed with 5%H2/Ar for 5 min after the treatment of part a. Parts b and d are the enlargements of the
partial region (1200−2200 cm−1) in parts a and c, respectively.

reason for this seeming contradiction, O-XPS and CO2-TPD 700 were calculated as 16%, 12%, and 10%, respectively, using
were performed. Two peaks were observed in the O-XPS the term Ovac/(Olat + Ovac).
spectra (Figure 7a) of the three catalysts: one peak near 529.8 According to the literature, oxygen vacancies on the surface
of catalysts could serve as active Lewis acidic sites and
eV was assigned to lattice oxygen (Olat), while the other peak
contributed significantly to the adsorption−desorption process
located at 531.8 eV was attributed to oxygen atoms in the of CO2.36,37 Therefore, CO2-TPD characterization was
vicinity of the oxygen vacancy (Ovac).35 The fractions of the conducted to obtain more information about the surface
oxygen vacancies in Ru/TiO2-A, Ru/TiO2-600, and Ru/TiO2- characteristics. As shown in Figure 7b, the area of the
14293 https://doi.org/10.1021/acssuschemeng.1c05565
ACS Sustainable Chem. Eng. 2021, 9, 14288−14296
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

desorption peak near 72 °C decreased dramatically with the through sequential hydrogenation after the dissociation of
increase of the rutile content of the support. In general, this formate, which was similar to the above conclusion from the
desorption peak (low-temperature region) was attributed to experimental observations.
the desorption of CO2 weakly adsorbed on the catalyst (e.g., On the basis of all of the characterization results and
adsorbed on surface hydroxyl groups).38,39 High-temperature literature reports, one can conclude that the rutile content in
calcination would not only lead to a sharp reduction in the Ru/TiO2 constituted an important contribution to the catalyst
specific surface area of the TiO2 support (Table S1, SI) but performance in low-temperature CO2 methanation in this
also lead to a dramatic drop in surface hydroxyl groups, work. The concentration of oxygen vacancies would be fairly
thereby resulting in a decrease in the adsorption amount of high if the rutile content was small, which could facilitate the
CO2.40 Moreover, the temperature of the desorption peak formation of *CO2.48,49 However, when the rutile content was
located in the high-temperature region also decreased from high, Ru species could be better dispersed on TiO2 due to the
316 °C for Ru/TiO2-A to 294 °C for Ru/TiO2-600, while no preferential growth, which was conducive to the stabilization of
obvious desorption peak was observed for Ru/TiO2-700. As the metal−oxide interface. Specifically, metallic Ru species
stated above, the oxygen vacancy concentration decreased could dissociate H2, and *HCOO was formed through the
significantly with the increase of the rutile content, which hydrogenation of CO2.50,51 Furthermore, the metal−oxide
caused the drastic decrease in the number of CO2 adsorption interface could enhance the binding ability of *HCOO to
sites (strong CO2 adsorption sites) on the support.41 Zhang et promote the dissociation of its C−O bond.50 After scission of
al.42 found that oxygen vacancies could promote the the C−O bonds, CHx species would be continuously
adsorption of CO2, and the order of CO2 adsorption ability hydrogenated and finally form CH4. When the rutile content
on Au/CeO2 was in accordance with the concentration of was about 15−30%, both factors were balanced to achieve the
oxygen vacancies, which was similar to the results in this highest reactivity for CO2 methanation. Kim et al.28 found that,
investigation. It should be stated that, during the activation for Ru on pure anatase and Ru on pure rutile, the amount of
process of CO2, the oxygen vacancies in the rutile phase and available Ru species was lower on the surface of the catalysts
anatase phase both are involved in the reaction due to the compared to Ru on the mixed-phase TiO2 support, thus
highly disordered and mixed status between them. In causing a decrease in reactivity. When it comes to P25 (80%
summary, the adsorption sites of CO2 on the catalyst anatase + 20% rutile), there is a relatively suitable interaction
significantly decreased with the increase of rutile content, strength between Ru species and P25, which led to a higher
which led to the decline in the amount of adsorbed CO2. As a catalytic activity in CO2 methanation. The rutile contents of
result, the reactivity of Ru/TiO2-700 was worse than that of P25 and Ru/TiO2-600 in this research are in the same range.
Ru/TiO2-600.
Furthermore, CO2 methanation was investigated by in situ
DRIFTS. The apparent activation energy (Ea) for Ru/TiO2-A,
■ CONCLUSION
In this work, catalysts with Ru supported on TiO2 with
Ru/TiO2-600, and Ru/TiO2-700 catalysts were 89, 86, and 86 different rutile contents were specifically prepared to
kJ/mol, respectively (Figure S9, SI). The similar Ea values investigate the effect of the support phase on the catalytic
suggested the presence of the same active species in these three performances of catalysts in low-temperature CO2 methana-
catalysts, meaning that the support phase did not affect the tion. It was found that Ru species grew preferentially on the
reaction mechanism.28,43 Consequently, Ru/TiO2-600 was rutile phase of the catalysts, and the deposited Ru and its
chosen to study the CO2 methanation reaction mechanism, metal−oxide interface displayed higher stability with increasing
and the characterization results are shown in Figure 8. Under rutile content. However, the characterization results of O-XPS
the reaction gas mixture, the signal intensities attributed to and CO2-TPD indicated that the concentration of oxygen
gaseous CH4 species (3016 and 1304 cm−1) increased vacancies (i.e., the adsorption sites for CO2) on the support
gradually as the reaction progressed at 300 °C (Figure 8a).44 decreased with the increase of rutile content. At a rutile
Simultaneously, the signal intensities ascribed to *HCOO content of 15−30% in Ru/TiO2-600, these two opposing
species (1508 and 1342 cm−1) exhibited a trend similar to that factors were balanced, resulting in the highest reactivity of this
of gaseous CH4 species (Figure 8b).45 When switching the gas catalyst among the investigated Ru/TiO2 materials. The in situ
from 15%CO2/45%H2/N2 to 5%H2/Ar at 300 °C, the peak DRIFTS analysis suggested *HCOO as the main intermediate
intensities of CH4 and *HCOO species decreased monotoni- species, leading to the conclusion that CO2 was first adsorbed
cally with increasing contact time (Figure 8c,d). More on oxygen vacancies and then *HCOO was formed by
specifically, the intensity of the peak assigned to Ru-bonded hydrogenation. Eventually, CH4 was generated through
carbonyl species (1982 cm−1) was extremely low and did not successive hydrogenation of CHx after the dissociation of the
change during these two experiments, indicating that the C−O bonds in *HCOO. In summary, this research studied the
adsorption of CO on Ru species was very weak.46 Meanwhile, effect of rutile content on reaction performance in detail and
CO-TPD curves also showed that CO had almost desorbed provided guidance to construct catalysts with high activity for
completely from the catalyst in He (99.999%) at room low-temperature CO2 methanation in the future.


temperature and no obvious desorption peak related to CO
adsorbed on the catalyst was observed in the heating process ASSOCIATED CONTENT
(Figure S10, SI), which also meant that the adsorption *
sı Supporting Information
strength of CO on the catalyst was very weak. Thus, one could
The Supporting Information is available free of charge at
conclude that CH4 might not be produced through the reverse
https://pubs.acs.org/doi/10.1021/acssuschemeng.1c05565.
water−gas shift reaction (RWGS) followed by CO hydro-
genation but that *HCOO might be the real intermediate The particle size and rutile content of each TiO2
species. Wang et al.47 discovered that CO2 methanation support, binding energy and valence state of Ru species,
followed the formate route on Ru/CeO2, and CH4 was formed the specific surface area (SBET) of samples, the stability
14294 https://doi.org/10.1021/acssuschemeng.1c05565
ACS Sustainable Chem. Eng. 2021, 9, 14288−14296
ACS Sustainable Chemistry & Engineering


pubs.acs.org/journal/ascecg Research Article

test of Ru/TiO2-600 catalyst, XRD patterns under ACKNOWLEDGMENTS


different conditions, TEM images and particle size This work was supported by the National Key Research and
distribution of catalysts after reduction and reaction, Development Program of China (2017YFA0206801) and the
XPS spectra of catalysts after calcination in air and National Natural Science Foundation of China (22071202,
reaction, apparent active energy test, and CO-TPD 21931009, 21721001, and 21773190).


curves of catalysts (PDF)


REFERENCES
AUTHOR INFORMATION (1) Vogt, C.; Monai, M.; Kramer, G. J.; Weckhuysen, B. M. The
renaissance of the Sabatier reaction and its applications on Earth and
Corresponding Authors in space. Nat. Catal. 2019, 2, 188−197.
Qin Kuang − State Key Laboratory for Physical Chemistry of (2) Li, W.; Wang, H.; Jiang, X.; Zhu, J.; Liu, Z.; Guo, X.; Song, C. A
Solid Surfaces, Collaborative Innovation Center of Chemistry short review of recent advances in CO2 hydrogenation to hydro-
for Energy Materials, Department of Chemistry, College of carbons over heterogeneous catalysts. RSC Adv. 2018, 8, 7651−7669.
Chemistry and Chemical Engineering, Xiamen University, (3) Davis, W.; Martín, M. Optimal year-round operation for
methane production from CO2 and water using wind and/or solar
Xiamen 361005, China; orcid.org/0000-0002-4111-
energy. J. Cleaner Prod. 2014, 80, 252−261.
291X; Email: zxxie@xmu.edu.cn (4) Yu, W.-Z.; Fu, X.-P.; Xu, K.; Ling, C.; Wang, W.-W.; Jia, C.-J.
Zhaoxiong Xie − State Key Laboratory for Physical Chemistry CO2 methanation catalyzed by a Fe-Co/Al2O3 catalyst. J. Environ.
of Solid Surfaces, Collaborative Innovation Center of Chem. Eng. 2021, 9, 105594.
Chemistry for Energy Materials, Department of Chemistry, (5) Hussain, I.; Jalil, A. A.; Hassan, N. S.; Hamid, M. Y. S. Recent
College of Chemistry and Chemical Engineering, Xiamen advances in catalytic systems for CO2 conversion to substitute natural
University, Xiamen 361005, China; Pen-Tung Sah Institute gas (SNG): Perspective and challenges. J. Energy Chem. 2021, 62,
of Micro-Nano Science and Technology, Xiamen University, 377−407.
Xiamen 361005, China; orcid.org/0000-0002-4225- (6) Vrijburg, W. L.; Moioli, E.; Chen, W.; Zhang, M.; Terlingen, B. J.
6536; Email: qkuang@xmu.edu.cn P.; Zijlstra, B.; Filot, I. A. W.; Züttel, A.; Pidko, E. A.; Hensen, E. J. M.
Efficient base-metal NiMn/TiO2 catalyst for CO2 methanation. ACS
Authors Catal. 2019, 9, 7823−7839.
(7) Du, Y.; Qin, C.; Xu, Y.; Xu, D.; Bai, J.; Ma, G.; Ding, M. Ni
Zhiying Zhao − State Key Laboratory for Physical Chemistry nanoparticles dispersed on oxygen vacancies-rich CeO2 nanoplates for
of Solid Surfaces, Collaborative Innovation Center of enhanced low-temperature CO2 methanation performance. Chem.
Chemistry for Energy Materials, Department of Chemistry, Eng. J. 2021, 418, 129402.
College of Chemistry and Chemical Engineering, Xiamen (8) Wierzbicki, D.; Baran, R.; Dębek, R.; Motak, M.; Gálvez, M. E.;
University, Xiamen 361005, China Grzybek, T.; Da Costa, P.; Glatzel, P. Examination of the influence of
Qiaorong Jiang − State Key Laboratory for Physical La promotion on Ni state in hydrotalcite-derived catalysts under CO2
Chemistry of Solid Surfaces, Collaborative Innovation Center methanation reaction conditions: Operando X-ray absorption and
of Chemistry for Energy Materials, Department of Chemistry, emission spectroscopy investigation. Appl. Catal., B 2018, 232, 409−
College of Chemistry and Chemical Engineering, Xiamen 419.
University, Xiamen 361005, China (9) Martin, N. M.; Velin, P.; Skoglundh, M.; Bauer, M.; Carlsson, P.-
A. Catalytic hydrogenation of CO2 to methane over supported Pd, Rh
Qiuxiang Wang − Instrumental Analysis Center, Huaqiao and Ni catalysts. Catal. Sci. Technol. 2017, 7, 1086−1094.
University, Xiamen 361021, China (10) Lin, S.; Hao, Z.; Shen, J.; Chang, X.; Huang, S.; Li, M.; Ma, X.
Mingzhi Wang − State Key Laboratory for Physical Chemistry Enhancing the CO2 methanation activity of Ni/CeO2 via activation
of Solid Surfaces, Collaborative Innovation Center of treatment-determined metal-support interaction. J. Energy Chem.
Chemistry for Energy Materials, Department of Chemistry, 2021, 59, 334−342.
College of Chemistry and Chemical Engineering, Xiamen (11) Moioli, E.; Züttel, A. A model-based comparison of Ru and Ni
University, Xiamen 361005, China catalysts for the Sabatier reaction. Sustainable Energy Fuels 2020, 4,
Jiachang Zuo − State Key Laboratory for Physical Chemistry 1396−1408.
of Solid Surfaces, Collaborative Innovation Center of (12) Sakpal, T.; Lefferts, L. Structure-dependent activity of CeO2
Chemistry for Energy Materials, Department of Chemistry, supported Ru catalysts for CO2 methanation. J. Catal. 2018, 367,
171−180.
College of Chemistry and Chemical Engineering, Xiamen (13) Guo, Y.; Mei, S.; Yuan, K.; Wang, D.-J.; Liu, H.-C.; Yan, C.-H.;
University, Xiamen 361005, China Zhang, Y.-W. Low-temperature CO2 methanation over CeO2-
Hanming Chen − State Key Laboratory for Physical supported Ru single atoms, nanoclusters, and nanoparticles
Chemistry of Solid Surfaces, Collaborative Innovation Center competitively tuned by strong metal-support interactions and H-
of Chemistry for Energy Materials, Department of Chemistry, spillover effect. ACS Catal. 2018, 8, 6203−6215.
College of Chemistry and Chemical Engineering, Xiamen (14) Yin, A.-X.; Liu, W.-C.; Ke, J.; Zhu, W.; Gu, J.; Zhang, Y.-W.;
University, Xiamen 361005, China Yan, C.-H. Ru nanocrystals with shape-dependent surface-enhanced
raman spectra and catalytic properties: Controlled synthesis and DFT
Complete contact information is available at: calculations. J. Am. Chem. Soc. 2012, 134, 20479−20489.
https://pubs.acs.org/10.1021/acssuschemeng.1c05565 (15) Kwak, J. H.; Kovarik, L.; Szanyi, J. Heterogeneous catalysis on
atomically dispersed supported metals: CO2 reduction on multifunc-
Author Contributions tional Pd catalysts. ACS Catal. 2013, 3, 2094−2100.
The manuscript was written through contributions of all (16) Zhao, Z.; Wang, M.; Ma, P.; Zheng, Y.; Chen, J.; Li, H.; Zhang,
authors. All authors have given approval to the final version of X.; Zheng, K.; Kuang, Q.; Xie, Z.-X. Atomically dispersed Pt/CeO2
the manuscript. catalyst with superior CO selectivity in reverse water gas shift
reaction. Appl. Catal., B 2021, 291, 120101.
Notes (17) Zhang, S.; Xia, Z.; Zhang, M.; Zou, Y.; Shen, H.; Li, J.; Chen,
The authors declare no competing financial interest. X.; Qu, Y. Boosting selective hydrogenation through hydrogen

14295 https://doi.org/10.1021/acssuschemeng.1c05565
ACS Sustainable Chem. Eng. 2021, 9, 14288−14296
ACS Sustainable Chemistry & Engineering pubs.acs.org/journal/ascecg Research Article

spillover on supported-metal catalysts at room temperature. Appl. for promoted visible-light water splitting. J. Am. Chem. Soc. 2014, 136,
Catal., B 2021, 297, 120418. 6826−6829.
(18) Rodriguez, J. A.; Liu, P.; Stacchiola, D. J.; Senanayake, S. D.; (36) Xin, C.; Hu, M.; Wang, K.; Wang, X. Significant enhancement
White, M. G.; Chen, J. G. Hydrogenation of CO2 to methanol: of photocatalytic reduction of CO2 with H2O over ZnO by the
Importance of metal-oxide and metal-carbide interfaces in the formation of basic zinc carbonate. Langmuir 2017, 33, 6667−6676.
activation of CO2. ACS Catal. 2015, 5, 6696−6706. (37) Zhang, S.; Xia, Z.; Zou, Y.; Cao, F.; Liu, Y.; Ma, Y.; Qu, Y.
(19) Chen, Q.; Chen, X.; Fang, M.; Chen, J.; Li, Y.; Xie, Z.; Kuang, Interfacial frustrated lewis pairs of CeO2 activate CO2 for selective
Q.; Zheng, L. Photo-induced Au-Pd alloying at TiO2 {101} facets tandem transformation of olefins and CO2 into cyclic carbonates. J.
enables robust CO2 photocatalytic reduction into hydrocarbon fuels. Am. Chem. Soc. 2019, 141, 11353−11357.
J. Mater. Chem. A 2019, 7, 1334−1340. (38) Zhong, C.; Guo, X.; Mao, D.; Wang, S.; Wu, G.; Lu, G. Effects
(20) Chen, Z.; Liang, L.; Yuan, H.; Liu, H.; Wu, P.; Fu, M.; Wu, J.; of alkaline-earth oxides on the performance of a CuO-ZrO2 catalyst
Chen, P.; Qiu, Y.; Ye, D.; Chen, L. Reciprocal regulation between for methanol synthesis via CO2 hydrogenation. RSC Adv. 2015, 5,
support defects and strong metal-support interactions for highly 52958−52965.
efficient reverse water gas shift reaction over Pt/TiO2 nanosheets (39) Gao, P.; Li, F.; Zhan, H.; Zhao, N.; Xiao, F.; Wei, W.; Zhong,
catalysts. Appl. Catal., B 2021, 298, 120507. L.; Wang, H.; Sun, Y. Influence of Zr on the performance of Cu/Zn/
(21) Wang, L.; Huang, G.; Zhang, L.; Lian, R.; Huang, J.; She, H.; Al/Zr catalysts via hydrotalcite-like precursors for CO2 hydrogenation
Liu, C.; Wang, Q. Construction of TiO2-covalent organic framework to methanol. J. Catal. 2013, 298, 51−60.
Z-Scheme hybrid through coordination bond for photocatalytic CO2 (40) Xie, Y. H.; Li, B.; Weng, W. Z.; Zheng, Y. P.; Zhu, K.-T.; Zhang,
N. W.; Huang, C. J.; Wan, H. L. Mechanistic aspects of formation of
conversion. J. Energy Chem. 2022, 64, 85−92.
sintering-resistant palladium nanoparticles over SiO2 prepared using
(22) Kattel, S.; Yan, B.; Chen, J. G.; Liu, P. CO2 hydrogenation on
Pd(acac)2 as precursor. Appl. Catal., A 2015, 504, 179−186.
Pt, Pt/SiO2 and Pt/TiO2: Importance of synergy between Pt and
(41) Liu, B.; Li, C.; Zhang, G.; Yao, X.; Chuang, S. S. C.; Li, Z.
oxide support. J. Catal. 2016, 343, 115−126.
Oxygen vacancy promoting dimethyl carbonate synthesis from CO2
(23) Lin, Y.; Zhu, Y.; Pan, X.; Bao, X. Modulating the methanation
and methanol over Zr-doped CeO2 nanorods. ACS Catal. 2018, 8,
activity of Ni by the crystal phase of TiO2. Catal. Sci. Technol. 2017, 7, 10446−10456.
2813−2818. (42) Zhang, X.; Wang, D.; Jing, M.; Liu, J.; Zhao, Z.; Xu, G.; Song,
(24) Li, W.; Zhang, G.; Jiang, X.; Liu, Y.; Zhu, J.; Ding, F.; Liu, Z.; W.; Wei, Y.; Sun, Y. Ordered mesoporous CeO2-supported Ag as an
Guo, X.; Song, C. CO2 Hydrogenation on unpromoted and M- effective catalyst for carboxylative coupling reaction using CO2.
promoted Co/TiO2 catalysts (M = Zr, K, Cs): Effects of crystal phase ChemCatChem 2019, 11, 2089−2098.
of supports and metal-support interaction on tuning product (43) Wang, F.; Li, C.; Zhang, X.; Wei, M.; Evans, D. G.; Duan, X.
distribution. ACS Catal. 2019, 9, 2739−2751. Catalytic behavior of supported Ru nanoparticles on the {100},
(25) Kim, A.; Debecker, D. P.; Devred, F.; Dubois, V.; Sanchez, C.; {110}, and {111} facet of CeO2. J. Catal. 2015, 329, 177−186.
Sassoye, C. CO2 methanation on Ru/TiO2 catalysts: On the effect of (44) Eckle, S.; Anfang, H.-G.; Behm, R. J. Reaction intermediates
mixing anatase and rutile TiO2 supports. Appl. Catal., B 2018, 220, and side products in the methanation of CO and CO2 over supported
615−625. Ru catalysts in H2-rich reformate gases. J. Phys. Chem. C 2011, 115,
(26) Li, X.; Lin, J.; Li, L.; Huang, Y.; Pan, X.; Collins, S. E.; Ren, Y.; 1361−1367.
Su, Y.; Kang, L.; Liu, X.; Zhou, Y.; Wang, H.; Wang, A.; Qiao, B.; (45) Arunajatesan, V.; Subramaniam, B.; Hutchenson, K. W.;
Wang, X.; Zhang, T. Controlling CO2 hydrogenation selectivity by Herkes, F. E. In situ FTIR investigations of reverse water gas shift
metal-support electron transfer under reaction conditions. Angew. reaction activity at supercritical conditions. Chem. Eng. Sci. 2007, 62,
Chem., Int. Ed. 2020, 59, 19983−19989. 5062−5069.
(27) Lin, Q.; Liu, X. Y.; Jiang, Y.; Wang, Y.; Huang, Y.; Zhang, T. (46) Panagiotopoulou, P.; Verykios, X. E. Mechanistic study of the
Crystal phase effects on the structure and performance of ruthenium selective methanation of CO over Ru/TiO2 catalysts: Effect of metal
nanoparticles for CO2 hydrogenation. Catal. Sci. Technol. 2014, 4, crystallite size on the nature of active surface species and reaction
2058−2063. pathways. J. Phys. Chem. C 2017, 121, 5058−5068.
(28) Kim, A.; Sanchez, C.; Patriarche, G.; Ersen, O.; Moldovan, S.; (47) Wang, F.; He, S.; Chen, H.; Wang, B.; Zheng, L.; Wei, M.;
Wisnet, A.; Sassoye, C.; Debecker, D. P. Selective CO2 methanation Evans, D. G.; Duan, X. Active site dependent reaction mechanism
on Ru/TiO2 catalysts: Unravelling the decisive role of the TiO2 over Ru/CeO2 catalyst toward CO2 methanation. J. Am. Chem. Soc.
support crystal structure. Catal. Sci. Technol. 2016, 6, 8117−8128. 2016, 138, 6298−6305.
(29) Ma, Y.; Wang, X.; Jia, Y.; Chen, X.; Han, H.; Li, C. Titanium (48) Huygh, S.; Bogaerts, A.; Neyts, E. C. How oxygen vacancies
dioxide-based nanomaterials for photocatalytic fuel generations. activate CO2 dissociation on TiO2 anatase (001). J. Phys. Chem. C
Chem. Rev. 2014, 114, 9987−10043. 2016, 120, 21659−21669.
(30) Padayachee, D.; Mahomed, A. S.; Singh, S.; Friedrich, H. B. (49) Ye, J.; Liu, C.; Mei, D.; Ge, Q. Active oxygen vacancy site for
Effect of the TiO2 anatase/rutile ratio and interface for the oxidative methanol synthesis from CO2 hydrogenation on In2O3(110): A DFT
activation of n-octane. ACS Catal. 2020, 10, 2211−2220. study. ACS Catal. 2013, 3, 1296−1306.
(31) Jagadale, T. C.; Takale, S. P.; Sonawane, R. S.; Joshi, H. M.; (50) Kattel, S.; Liu, P.; Chen, J. G. Tuning selectivity of CO2
Patil, S. I.; Kale, B. B.; Ogale, S. B. N-doped TiO2 nanoparticle based hydrogenation reactions at the metal/oxide interface. J. Am. Chem.
visible light photocatalyst by modified peroxide sol-gel method. J. Soc. 2017, 139, 9739−9754.
Phys. Chem. C 2008, 112, 14595−14602. (51) Wang, W.; Wang, S.; Ma, X.; Gong, J. Recent advances in
(32) Gerhard, E.; Helmuth, K.; Ferdi, S.; Jens, W. Handbook of catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40,
Heterogeneous Catalysis; Wiley, 2008. 3703−3727.
(33) Folkesson, B. ESCA studies on the charge distribution in some
dinitrogen complexes of rhenium, iridium, ruthenium, and osmium.
Acta Chem. Scand. 1973, 27, 287−302.
(34) Chen, S.; Abdel-Mageed, A. M.; Gauckler, C.; Olesen, S. E.;
Chorkendorff, I.; Behm, R. J. Selective CO methanation on
isostructural Ru nanocatalysts: The role of support effects. J. Catal.
2019, 373, 103−115.
(35) Lei, F.; Sun, Y.; Liu, K.; Gao, S.; Liang, L.; Pan, B.; Xie, Y.
Oxygen vacancies confined in ultrathin indium oxide porous sheets

14296 https://doi.org/10.1021/acssuschemeng.1c05565
ACS Sustainable Chem. Eng. 2021, 9, 14288−14296

You might also like