Boosting CO2 methanation activity on Ru_TiO2 catalysts by exposing (001) facets of anatase TiO2

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of CO₂ Utilization 33 (2019) 242–252

Contents lists available at ScienceDirect

Journal of CO2 Utilization


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

Boosting CO2 methanation activity on Ru/TiO2 catalysts by exposing (001) T


facets of anatase TiO2

Shanshan Chaia, Yong Mena, , Jinguo Wanga, Shuang Liua, Qiaoling Songa, Wei Ana,
⁎⁎
Gunther Kolbb,c,
a
College of Chemistry and Chemical Engineering, Shanghai University of Engineering Science, Shanghai, 201620, PR China
b
Department of Chemical Engineering and Chemistry, Eindhoven University of Technology, P.O Box 513, 5600 MB, Eindhoven, the Netherlands
c
Energy and Chemical Technology Division, Fraunhofer IMM, Carl.Zeiss-Str.18-20, D-55129, Mainz, Germany

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

Keywords: Catalytic reduction of CO2 to methane has been recognized as one of the most important strategic reactions to
Ru nanoparticles produce highly valuable chemicals and for storage of renewable energy (power-to-gas). This study demonstrates
TiO2 nanocrystal a crystal facet-dependent catalytic reduction of CO2 to CH4 on ruthenium nanoparticles deposited over TiO2
Exposed (001) facets nanocrystal with exposed (001) and (101) facets. Compared with TiO2 (101) nanocrystal with the same Ru
CO2 methanation
loading, Ru nanoparticles supported by TiO2 (001) exhibited a highly enhanced CO2 conversion rate and sig-
synergistic effect
nificantly improved methanation reactivity, along with a considerable durability. The physicochemical prop-
erties of the samples were characterized by HR-TEM, XRD, BET, FT-IR, UV–vis, Raman, ICP, and temperature-
programmed reaction with CO2, and CO2 + H2. In view of the difference in catalytic activity and the physi-
cochemical properties of the supported Ru/TiO2 catalysts, the results uncovered that the nature of the catalyst
support of Ru/TiO2 strongly affected the dispersion of Ru species and the synergistic effect between Ru and
underlying TiO2 supporting materials due to the strong metal-support interaction, and thus affected their cap-
ability to activate CO2 and determined the catalytic activity for CO2 methanation. Our work demonstrated the
important role of the exposed crystal facets of the underlying support and the resulting Ru particle which had a
large effect on CO2 methanation.

1. Introduction to its worldwide reserves and its economic and environmental ad-
vantages over coal [13]. The methanation of CO2 is regarded as a
As the major contributors for the greenhouse effect [1–3] and in- simple and highly efficient route towards methane. CO2 conversion
exhaustible source of cheap carbon and oxygen resources [4,5], the with H2 is highly exothermic. Sabatier and Senderens first discovered
utilization of carbon dioxide has become one of the major strategic transition metals catalysts for methanation reaction in the early 20th
research topics worldwide [2,6]. The hydrogenation of CO2 can not century. Ni is the most widely applied and investigated catalysts be-
only mitigate the global greenhouse phenomenon but also address the cause of its high reactivity and relatively low cost [14–19]. However,
growing energy demand. The photolysis and electrolysis of water using Ni-based catalysts are also vulnerable to deactivation by coke formation
renewable energy could provide the hydrogen required for CO2 con- through either CH4 decomposition or via CO disproportionation (Bou-
version [7,8]. Therefore, to reduce the reliance on fossil fuels and keep douard reaction) and through sulphur poisoning. Noble metals in-
carbon balance in nature, the use of carbon dioxide as a carbon source cluding Ru [20–25], Pd [26,27], Pt [28–30], and Rh [31,32], are also
for valuable products such as methane, methanol, ethanol, dimethyl active for CO2 methanation. Among these noble metals, Ru is the most
ether, formic acid and higher hydrocarbons by hydrogenation of CO2 active catalyst in a wide range of operation temperatures.
has been extensively studied in the past decades [6,9–12]. The activity and selectivity of supported Ru-based catalysts are
Methane has been considered to become a major source of energy strongly affected by the preparation method, the amount of metal
and feed-stock for chemicals production after the petroleum age owing loading, the size of the dispersed metal particles, the interactions


Corresponding author.
⁎⁎
Corresponding author at: Department of Chemical Engineering and Chemistry, Eindhoven University of Technology, P.O Box 513, 5600 MB, Eindhoven, the
Netherlands.
E-mail addresses: men@sues.edu.cn (Y. Men), Gunther.kolb@imm.fraunhofer.de (G. Kolb).

https://doi.org/10.1016/j.jcou.2019.05.031
Received 25 March 2019; Received in revised form 25 May 2019; Accepted 26 May 2019
Available online 03 June 2019
2212-9820/ © 2019 Elsevier Ltd. All rights reserved.
S. Chai, et al. Journal of CO₂ Utilization 33 (2019) 242–252

between the support and the active metallic species, and the nature of of TiO2 on the structure and catalytic properties of the Ru nanoparticles
the support material. Abe et al. have developed a dry technique named for CO2 methanation have not yet been well established. In this work,
polygonal barrel-sputtering for modifying the surfaces of powder ma- single phase anatase became our choice of catalyst support to unravel
terials to deposit Ru nanoparticles onto TiO2 with narrow particle size the role of different crystal facets played in the CO2 methanation and
distribution without the need of any heating which can cause nano- Ru/TiO2 catalysts by exposing (001) and (101) facets of TiO2 were
particle sintering [33]. They reported that the turnover number de- prepared and studied for CO2 methanation. By using these Ru/TiO2
termined for the CO2 methanation reaction decreased with increasing catalysts with specifically exposed (001) and (101) crystal facets, we
size of Ru nanoparticles up to 6 nm and concluded that a low Ru par- are able to investigate and understand the role played by (001) and
ticle size is essential for a high CO2 methanation reactivity for the TiO2- (101) crystal-facet for CO2 methanation, and to uncover the influence of
supported Ru nanoparticle catalyst. Sassoye et al. have developed crystal facets of TiO2 support on the structure and catalytic properties
highly stable and well defined monodispersed aqueous colloidal sus- of the Ru nanoparticles for CO2 methanation. In this study, Ru deco-
pensions of small ruthenium nanoparticles by adding a certain amount rated over the TiO2 (001) facets exhibited excellent activity and dur-
of hydrogen peroxide. Highly dispersed ruthenium nanoparticles have ability for CO2 methanation compared to the TiO2 with exposed (101)
been proposed to be highly effective for the production of methane facets. The highly enhanced catalytic activity of Ru species loaded on
from CO2 [34]. However, Kwak et al. reported that the activity of Ru/ TiO2 (001) facets was ascribed to the nature of (001) facets with high
Al2O3 catalysts for CO2 reduction is determined by the size of the metal surface energy and abundant oxygen vacancies and the synergistic ef-
particles, and concluded that large metal clusters are favorable for CH4 fect between Ru and TiO2. The method of preparation of catalysts was
formation [22,27]. Several excellent reviews describe that the catalysts selected based on the research experience of crystal facet effect over
performance and stability are closely related to their particle size TiO2 in our group and the previous studies by other groups [40,55,56].
[31,35–38]. It has been recognized also that the chemical nature of the
support or the metal-support interactions can have a marked impact on 2. Experimental section
the catalytic activity and selectivity [39]. Reducible metal oxides in-
cluding TiO2, ZrO2 and CeO2 have been widely used as support mate- 2.1. Catalyst preparation
rials [20,21,40,41]. Among all the investigated Ru/oxide catalysts for
CO2 methanation, TiO2 was regarded as the most efficient support for TiO2 nanocrystal with exposed (001) and (101) facets were pre-
CO2 methanation [10]. The crystal phase of the TiO2 support plays an pared by solvothermal hydrolysis according to previously published
important role that affects the structure and catalytic properties of the work by the authors of the current paper [55]. To prepare TiO2 nano-
Ru nanoparticles, dictates the morphology of Ru species. Kim et al. crystals with exposed (001) facets, tetrabutyl titanate (2.6 M, 50 mL,
demonstrated that the catalytic activity of CO2 methanation was CP, with a concentration of 98 wt.%) and HF (3.2 M, 8 mL, AR, with a
markedly affected by the crystallinity of the TiO2 support, and stated concentration of 40 wt.%) mixed and stirred at room temperature, then
that the RuO2 decorated on the mixing of anatase and rutile particles transferred into a 100 ml Teflon-lined autoclave and sealed. The auto-
(P25) shows a better performance for CO2 methanation, followed by clave was then placed in an electric oven and kept at 180 °C for 24 h.
single phase anatase and rutile-supported catalysts [42]. Then the precipitates were collected by centrifugation, washed with
Tailoring the crystal facets of catalyst support is important for distilled water several times and dried overnight at 80 °C, and the ob-
modulating the activity of high-performance catalysts with well-defined tained sample was denoted as T001 (all reagents are from Aladdin).
surface structures [43–46]. Extensive experimental and theoretical To prepare TiO2 nanocrystal with exposed (101) facets, tetrabutyl
studies have both demonstrated that TiO2 with exposed (001) facets are titanate (0.6 M, 15 mL, CP, with a concentration of 98 wt.%) mixed with
preferred than those with the (101) facets for being used as catalysts acetic acid (13.3 M, 60 mL, AR, with a concentration of 36 wt.%) at
and supports in many applications of catalysis, photocatalysis and room temperature and also transferred into a 100 ml Teflon-lined au-
electroctalysis [44,47]. For instance, Wang et al. reported that the ex- toclave, sealed and placed in an electric oven at 200 °C for 24 h. The
posed (001) facets possesses high surface energy and abundant oxygen precipitates were collected by centrifugation, washed with distilled
vacancies, thus favored the photocatalytic oxidation of aromatic alco- water several times and dried overnight at 80 °C, and the obtained
hols [48]. A DFT study on formic acid adsorption by Gong et al. showed sample was denoted as T101 (all reagents are from Aladdin).
anatase (001) possess O2c atoms with higher energies compared to The Ru/TiO2 catalysts were prepared as previously described
(101) containing the bulk O3c, demonstrating that the high reactivity of [40,56], and 2.5 wt.% Ru was added as nominal loading. In brief, they
anatase (001) is associated with low-coordinated O2c atoms [49]. A were obtained by a drop wise addition of 15%v/v H2O2 (AR, 30 wt.% in
combined experimental and DFT study of the CO2 capture by TiO2 H2O, from Aladdin) diluted in H2O into 0.011 M RuCl3·xH2O (x = 3–5,
anatase surfaces carried out by Mino et al. [50] revealed that the for- GR, from Sinopharm Chemical Reagent) dissolved in H2O so that the
mation of carbonate and bicarbonate species occurs preferentially on final concentration of Ru ≈ 0.007 M. The solution was heated at 95 °C
the anatase (001) surface, and the higher reactivity of this surface is for 2 h. Once cooled to room temperature, an appropriate amount of
related to a stronger (Lewis and Brønsted) basicity of the surface oxygen TiO2 powder was added to the colloidal suspension of RuO2 nano-
sites. Cao et al. reported that the enhanced photoactivity for CO2 me- particles to yield 2.5 wt.% of Ru in the final catalyst. The mixture was
thanation is ascribed to the increasing ratio of exposed TiO2 nano- put in an oven at 120 °C overnight then calcined at 400 °C for 4 h in
crystals facets from 5% to 51% (001) and the creation of surface het- static air and washed several times with water to remove the chloride,
erojunction between (001) and (101) crystal facets [51]. Moreover, Jin finally dried at 120 °C for 10 h. They were denoted as RT001 and
et al. reported that the higher photocatalytic CO2 reduction activity of RT101, respectively.
Pt nanoparticles over TiO2-001 was ascribed to the enhanced photo-
induced carrier separation efficiency than that on TiO2-010, which can 2.2. Catalysts characterization
be further correlated with the different CO2 adsorption capability on
anatase TiO2-001 and TiO2-010 facets due to their different surface The samples were characterized by using high-resonance transmis-
electronic and atomic structures [52]. sion electron microscopy (HRTEM, JEOL JEM-2100). Prior to TEM
The crystal phases and crystal facets of oxide support are two im- measurements, a small amount of powder sample was ultrasonically
portant factors in determining the reactivity of metal-supported cata- dispersed in an aqueous solution, and then the solution was dropped
lysts. Despite the considerable progress achieved about the support onto the copper network. The crystalline phases of the samples were
effect, in particular, the impact of crystal phases on CO2 methanation identified by X-ray powder diffraction (XRD) using a diffractometer (D/
over Ru/TiO2 catalysts so far [33,53,54], the influence of crystal facets Max-r B) applying Cu-Kα radiation (λ =1.54056 Å) over the scan range

243
S. Chai, et al. Journal of CO₂ Utilization 33 (2019) 242–252

of 2θ between 20°-80°. The mean crystallite size (D) of the samples was 2.3. Catalysts activity measurements
calculated via the Scherrer equation:
The activity of the catalyst was measured in a continuous flow fixed
Kλ bed reactor and the reaction temperature was measured in the middle
D=
β1 cosθ (1) of the catalyst bed with a K-type thermocouple placed within a quartz
2
capillary well (see Fig. S1). 200 mg catalyst was placed into a quartz
The BET surface area (SBET), pore volume (VP) and pore diameter tube reactor with an internal diameter of 6 mm. Prior to the test, the
(DP) were measured on a Micromeritics ASAP 2460 instrument at 77 K. catalysts were pretreated at 500 °C in H2 (99.999%, 30 ml/min) for
Prior to measurement, the samples were degassed at 100 °C with 1 h in 1.0 h. The reaction was conducted from 150 °C to 400 °C at atmospheric
flowing N2 (99.999%). An inductive coupled plasma emission spectro- pressure and heated with a temperature rate of 10 °C/min, feed gases
meter (ICP, Varian, VISTAMPXICP) was used to test the actual loading were consisted of 10% CO2 (99.999%), 40% H2 and 50% N2 (99.999%)
of Ru nanoparticles. Fourier-transform infrared (FT-IR) spectra of the at a flow rate of 20 ml/min and were analyzed via gas chromatography.
samples were collected by using a Nicolet 380 spectrometer. Prior to The product analysis in this work was performed on Shimazu GC-2014C
the measurement, the samples were pressed into a thin self-supported online-GC equipped with one TCD and one FID detector. In this work,
wafer and placed in the IR cell. FT-IR spectra of the catalysts were re- packed PQ column/C13X molecular sieve column and TCD detector are
corded in air within the range of 450–4000 cm−1. UV–vis diffuse re- used to separate and detect simple gases. The Rt-Q-BOND PLOT column
flectance spectra (UV–vis DRS) were recorded in the absorbance mode and FID are used to separate and detect hydrocarbons and oxygenates.
at room temperature within the range of 200–800 nm. During the Product analysis was taken by averaging two measurements within a
measurement, BaSO4 (from Aladdin) was used as background material TOS of 1.0 h at each temperature and a temperature step of 25 °C.
in a UV-2450 spectrophotometer (Shimadzu, Japan). And Raman And to test the effect of H2O, the water (under condition of 5 vol.%
spectra were collected on NEXUS-470, MC-2530 and Dilor Super H2O) was injected continuously at a rate of 0.9 cm3·h−1 by syringe
LabRam II respectively. pump into a vaporizer. For the test of CO methanation over the RT001
Hydrogen temperature-programmed reduction (H2-TPR) was car- sample, the feed gases were consisted of 10% CO (99.999%), 40% H2
ried out on a Micromeritics Autochem II 2920 instrument equipped and 50% N2 at a flow rate of 20 ml/min. As for the effect of space
with a thermal conductivity detector (TCD) to quantify the consumed velocity test, the RT001 catalyst was tested on 250 °C with a total flow
H2. The catalysts (50 mg) was pretreated in Ar (99.999%) atmosphere rate of 20 ml/min (10 vol.% CO2, 40 vol.% H2 diluted in N2), and the
at 200 °C for 1.0 h to remove adsorbed water, and then cooled to 50 °C space velocity was varied by changing the amount of catalyst. All of the
in Ar, followed by heating to 400 °C at a heating rate of 10 °C/min in a separated gases were detected by a thermal conductivity detector
flow of 10% H2/Ar. (TCD). All the gases used in the activity measurements were purchased
The Ru dispersion was determined through H2-TPD on a from Shanghai Shenkai Gas Technology Co. Ltd.
Micromeritics Autochem II2920 instrument. Prior to the TPD tests, The relevant test results were calculated by the following equations:
50 mg of the catalyst was reduced for an hour in a flow of 10% H2/Ar at CO2 (in) − CO2 (out)
500 °C, and purged with Ar at 500 °C for 30 min, then cooled to 50 °C in CO2 Conversion (%) = × 100
CO2 (in) (3)
Ar. Finally the adsorption of H2 took place and then Ar was introduced
to perform the TCD measurement with a temperature ramp of 10 °C/ CH 4 (out)
CH 4 selectivity (%) = × 100
min to 900 °C. The dispersion of Ru was calculated based on the volume CH 4 (out) + CO(out) (4)
of chemisorbed H2 using the following simplified equation:
CO(out)
CO selectivity (%) = × 100
2 × Vad × Mmetal × SF CH 4 (out) + CO(out) (5)
D% =
p × Vm × dr (2)
Y ield CH4 (%) = X CO2 ×SCH4 × 100 (6)
where P denotes the weight fraction of Ru in the sample as determined CO2 (in) × X CO2 ×SCH4
by ICP; Vm is the molar volume of H2 at standard temperature and CH 4 formation rate(μmol·S−1·g−1 ) =
m (7)
pressure(STP); dr is the reduction degree of Ru; Vad (mL·g−1) is the
volume of chemisorbed H2 at STP in the TPD procedure;M is the Where CO2 (in) and CO2 (out) are the molar flows of the inlet and outlet
molecular weight of Ru (101.07 g·mol−1); and SF is the stoichiometric carbon dioxide, respectively, CH4(out) and CO(out) are the molar flows
factor (Ru:H molar ratio in the chemisorption) which was assumed to of the outlet methane and carbon monoxide.
be 1. The turnover frequency (TOF) was calculated based on the formula:
Temperature programmed desorption of carbon dioxide was per- CO2 (in) × X CO2
TOF=
formed on an Micromeritics Autochem II 2920 instrument coupled with m× M
P
×D%
metal (8)
a Hiden HPR20 mass spectrometer. Before the tests 50 mg of the cata-
lysts were reduced in flowing 10% H2/Ar at 500 °C for 1.0 h. After Where X CO2 is CO2 conversion, m is the catalyst mass, P denotes the
cooling down to 50 °C in an Ar stream, then 5% CO2/He was periodi- weight of Ru in the sample as determined by ICP; Mmetal is the molar
cally injected into the sample until saturation. The sample was then weight of Ru; D% is the Ru dispersion, which was calculated by the H2-
purged with He (99.999%) at room temperature for 60 min, and heated TPD.
with a temperature rate of 10 °C/min to 900 °C in He while the effluent
gases were monitored by a mass spectrometer with special attention to 3. Results and discussion
the mass spectrometer m/e values of 15 (CH4), 44(CO2), 28 (CO) and 18
(H2O), respectively. 3.1. Morphology and physical characteristics of the catalysts
Temperature-programmed surface reaction (TPSR) measurements of
samples were performed in the same manner as the temperature pro- Fig. 1 shows the XRD patterns of the TiO2 support before and after
grammed desorption of carbon dioxide described above, except that the loading of Ru. It can be seen that all the samples showed a typical
sample was heated under 10% H2/Ar to 900 °C instead of He. All the anatase phase with high crystallization degree, corresponding to 2θ of
gases used in the experimental were purchased from Shanghai Shenkai 25.8°, 37.7°, 47.8°, 54.0°,55.0° and 62.7° indicative of (101), (004),
Gas Technology Co. Ltd. (200), (105), (211) and (204) diffractions respectively (JCPDS file card
No. 12-1272) [48,55]. The phase of RuOx and Ru can hardly be

244
S. Chai, et al. Journal of CO₂ Utilization 33 (2019) 242–252

summarized in Table 1. After loading the Ru nanoparticles, the SBET of


RT001 and RT101 declined to 67.9 m2·g−1 and 69.5 m2·g−1, respec-
tively. Moreover, loading Ru over the T001 resulted in an increase of
the pore diameter but a decreasing pore diameter was found for the
T101 sample, while the pore size distribution of all the samples are
relatively narrow and below 25 nm.
Fig. S2 presents the FT-IR spectra of all the samples in the region of
450–4000 cm−1, which were similar to those reported in literature
[58]. The broad absorption band at 3438 cm−1 in the high energy re-
gion was ascribed to the stretching vibration of hydroxyl group of the
water molecule physically adsorbed on the sample surface [59]. The
bands at 1636 cm−1 can be assigned to the bending vibration of the
water molecules. The band located at 1460 cm−1 is an indicative of the
presence of the absorption peaks of COH, which were generated by
insertion of some organic moieties into TiO2 during the course of sol-
vothermal reaction. The bands at around 480 cm−1 in the spectra of all
the samples are unambiguously assigned to the characteristic adsorp-
tion band of Ti-O, while the bands tended to be weakened after loading
of Ru species, suggesting a perturbation of the TiO2 framework upon
Fig. 1. XRD patterns of all the samples. interaction with Ru. Fig. 4 depicts the UV–vis DRS spectra of all the
samples. The pure TiO2 samples T001 and T101 both show strong ab-
Table 1 sorption at around 385 nm (λ < 400 nm) in the UV region. The spectra
Structural parameters of the catalysts. around 200–300 nm show clearly that the absorption capability of T001
is stronger than that of T101. Upon introducing Ru nanoparticles over
Sample SBET/m2·g−1 Vp/cm3·g−1 Dp/nm Crystal sizea/ Exposed (001)
nm (%) the two supports, the band gaps were significantly altered. The Ru
loaded samples showed a stronger absorption capability in the range of
T001 110.1 0.35 10.9 12.9 80.0 350–800 nm but slightly weaker absorption in the range of 200–350 nm
T101 92.6 0.25 10.4 18.9 8.0
than their respective pure supports. It was also noted that the loading of
RT001 67.9 0.29 15.3 13.1 80.0
RT101 69.5 0.17 7.4 19.2 8.0 Ru nanoparticles didn’t change the sequence of the absorption cap-
ability on TiO2. The difference in the absorption could be related to the
a
As calculated by the Scherrer equation. strong metal-support interaction, while the specific surface area of
RT001 and RT101 are very similar.
detected by the XRD, probably owing to the high degree of dispersion of The crystalline phases and surface oxygen vacancies were further
the Ru species on the TiO2 surface. The octahedron TiO2 mainly ex- determined using Raman spectroscopy. Parker et al. have demonstrated
posed (101) facets, whereas (001) facets were the mainly exposed facets that the shifting and broadening of the Raman spectral peaks were
over decahedron TiO2, as the (001) facets were formed by the de- ascribed to the oxygen stoichiometry of the material by using a series of
formation growth of anatase TiO2 crystal from octahedron to decahe- heat treatment experiments [60]. They also measured the oxygen intake
dron. And the exposed (001) facets was grown when the up and down of the nanophase TiO2 samples by using the thermogravimetric analysis
crystal growth was retarded. (See Scheme S1 (a)). According to Scherrer (TGA). Their results revealed that the Raman spectra are relevant to the
equation, the increased half-height width of (004) is corresponding to oxygen stoichiometry of the samples, (Eg) vibrational mode and the
“h” in Scheme S1 (a), implying the increased percentage of exposed peak position and full width at half maximum of the anatase
(001) facets. As for the increased intensity of (101) peak is corre- “143 cm−1” present the O/Ti ratio of sample. The positive shift of
sponding to the “m” in Scheme S1 (a) [57], Scheme S1 (a) is a cutway Raman peak position toward a higher wavenumber is the experimental
view of Scheme S1 (b), the angle between (101) and (001) facets was evidence for the increased surface oxygen vacancies in TiO2 [61,62]. It
68.3°, thus the percentage of (001) facets could be calculated according can be seen from Fig. 5, that the signals of all the samples recorded 5
to the Scherrer equation (see details in supplementary information). distinctive peaks at 144(Eg) cm−1, 197(Eg) cm−1, 397(B1g) cm−1,
The mean crystallite size (D) and the percentage of exposed (001) facets 519(A1g) cm−1 and 640(Eg) cm−1 which are all corresponding to the
of the samples were calculated and summarized in Table 1. Only 8% of anatase phase of TiO2. This result is in good agreement with the XRD
TiO2(001) facets were exposed in the T101 and RT101, whereas the patterns. From 110 to 180 cm−1 in the magnified Raman spectra,
T001 and RT001 exposed 80% of (001) facets. compared to T101, the T001 sample showed a positive shift of the
Fig. 2 presents the HRTEM and FFT images of all the samples. Ob- principal peak by about 2.0 cm−1. This Raman positive shift indicates
viously, T001 and RT001 both show a highly truncated bipyramidal that the number of surface oxygen vacancies in the T001 sample is
morphology, and the flat square surfaces are identified as (001) facets higher than that of the T101 sample.
with an average length around 25 nm and thickness about 5.0 nm. For Besides, after introducing Ru species, the RT001 sample also has a
the T101 and RT101 samples, they showed an octahedron morphology positive shift when compared with the RT101 sample, indicating the
and the isosceles trapezoidal surfaces are (101) facets with an average RT001 also possesses an abundant surface oxygen vacancies on (001)
size of 20 nm. Besides, from the HRTEM and FFT images, one can see facets. Analogous results reported previously have been attributed to
that all the samples were presented as well-defined anatase TiO2 single the increased number of surface oxygen vacancies on (001) facets [62],
crystals. The average particle size of the ruthenium nanoparticles of the This is mainly because the Ti atom in (001) facets is 5-fold coordinated,
T001 sample was determined to be 1.35 nm which is more than double while it is coordinated with either 5 or 6 oxygen atoms in (101) facets.
the value of 0.56 nm which was found for the T101 sample. Comparing with the (101) facets, the (001) facets had much more
All the samples displayed a typical type IV of N2 adsorption-deso- abundant surface oxygen vacancies facets due to the lower coordination
rption isotherms with H3 hysteresis loop in the relative pressure (P/P0) with Ti atoms (See Fig. S3). In accordance with the XRD results, T001
range from 0.8 to 1.0 in Fig. 3. These samples possessed a typical me- and RT001 both exposed a higher extent of (001) facets than T101 and
soporous structure due to the agglomeration of nanoparticles and the RT101, and consequently the T001 and RT001 could provide more
SBET, VP and DP values were calculated by BET equation and oxygen vacancies on (001) facets which serve as active sites for CO2

245
S. Chai, et al. Journal of CO₂ Utilization 33 (2019) 242–252

Fig. 2. TEM images of all the samples; the insets are the corresponding HRTEM and Fast-Fourier transform (FFT) images, respectively.

activation. showed the similar up-shifting of hydrogen reduction peaks over RT001
The strong metal-support interaction (SMSI) effect was originally catalyst, indicating the enhanced interaction of Ru and TiO2 (001)
proposed by Tauster et al. to describe the drastic changes in the phy- support.
sical and chemical properties of group VIII noble metals particles dis-
persed on reducible oxides [63,64]. And the strong interaction between
3.2. Catalytic activities of all the samples
noble metal and support was generated due to the inhibited H2 che-
misorption over catalysts and the extent of chemisorption suppression
Fig. 7(a) compares the catalytic performance of all the samples.
usually taken as an indicator of the extent of metal-support interaction
During the experiments of CO2 methanation, CO and CH4 were the only
[65]. However, the strong metal-support interaction has also been often
major products. Apparently, without addition of Ru nanoparticles to the
used in the recent catalysis literatures to describe the synergistic effect
two supports, the bare catalyst supports of T001 and T101 both show
between metal and support that could enhance catalytic activity in a
very low catalytic activity for the CO2 methanation during the whole
concerted manner [10,66,67]. In the literatures, this synergistic effect
process, the conversion of CO2 on the T001 is slightly higher than the
can be experimentally evidenced by the enhanced activity and some
T101, and the selectivity of CH4 shows very similar result. The two
specific characterization techniques e.g. up-shifted hydrogen reduction
supports only showed about 2% selectivity for CH4, while nearly 98%
peaks in H2-TPR [53,66–68]. H2-TPR results of all the samples in Fig. 6
selectivity for CO. After loading Ru nanoparticles, the activity of

Fig. 3. N2 adsorption-desorption isotherms (a) and pore size distribution curves (b) of the samples.

246
S. Chai, et al. Journal of CO₂ Utilization 33 (2019) 242–252

increased with the reaction temperature and become substantial at


300 °C. By comparison, the RT001 catalyst exhibited a much higher
activity than the RT101 for CO2 methanation. The yield of CH4 over the
RT001 catalyst increased with the reaction temperature, and ap-
proached the equilibrium conversion at 350 °C. A study of the effect of
H2O showed that introducing water into the feed had an inhibiting
effect on the reaction. RT001 still exhibited a superior catalytic activity
for CO2 methanation in the presence of H2O.
The reaction rates for CH4 are presented in Fig. 7(d), the CH4 pro-
duction rates of the catalysts at 300 °C are shown in Table 2. The CH4
production rate over the RT001 catalysts was 5.12 μmol·s−1·g−1 , which
is much higher than the value of 3.39 μmol·s−1·g−1 over the RT101
catalyst, illustrating that the T001 is a superior support for Ru for the
hydrogenation of CO2 to CH4.
In order to deeply investigate the intrinsic difference between the
RT001 and RT101 catalysts, a kinetic study was performed under the
kinetically controlled regime which was proved by the Weisz-Prater
−r ′A(obs) ρc R2
criterion CWP = De CAs
< 1 (see more details in supplementary in-
Fig. 4. UV–Vis diffuse reflection spectra (DRS) of all the samples. formation). Fig. 8(a) displays the Arrhenius plots of the two catalysts,
providing the apparent activation energies for CO2 methanation and
showing reasonable linearity in the temperature range under in-
vestigation (150–190 °C). As listed in Table 2, the apparent activation
energy for the RT001 and RT101 samples is 69.8 KJ/mol and 74.6 KJ/
mol, respectively, indicating that the intrinsic activity of CO2 metha-
nation on the RT001 and RT101 catalysts are largely affected by the
exposed TiO2 crystal facets and their associated different Ru particle
sizes. The different methanation activities may be explained by the
different amounts of active sites originating from different dispersion of
Ru on the carrier materials and a different number of surface oxygen
vacancies.
The stability of RT001 catalyst was evaluated under atmospheric
pressure at a WHSV = 6 L·h−1·g−1 and a reaction temperature of
325 °C. As shown in Fig. 8(b), all the CO2 conversion was always above
80% without any apparent deactivation. The CH4 selectivity remained
100%, and no selectivity towards CO was detected for the duration of
the test. The RT001 catalyst demonstrated therefore a very stable per-
formance for 50 h. Fig. S4 showed that the exposed (001) crystal facets
were still present after 50 h of stability test. In addition, the TEM inset
in Fig. S4 demonstrated that after 50 h stability test the RT001 sample
Fig. 5. Raman spectra of all the samples; the inset shows the magnified Raman
spectra from 110 to 180 cm−1.
remained the original truncated bipyramidal morphology, and the Ru
nanoparticles are well dispersed on T001 with the average size of Ru
nearly same as that on the fresh sample. H2-TPD measurements showed
that the Ru dispersion over the used RT001 is 31.1%, similar to the one
on the fresh sample (32.0%). Moreover, the stable crystalline structure
was also confirmed by the unchanged XRD patterns of fresh and used
RT001 samples. Therefore, the excellent stability of RT001 catalyst is
assumed to be closely associated with the strong anchoring of the Ru
nanoparticles on the (001) crystal facets of the TiO2 nanocrystals.
Fig. 9 displays the catalytic activity at 250 °C over the RT001 cat-
alysts at different space velocities. It was found that CH4 selectivity
increased with decreasing space velocity at the expense of CO. Un-
expected high CO selectivity was found on RT001 at a low CO2 con-
version, and the CO selectivity increased with increasing space velo-
cities. This finding strongly suggests that CO is a primary product that
undergoes secondary reactions through successive CO hydrogenation
over the Ru/TiO2 catalysts. In the literatures, the reaction mechanisms
proposed for CO2 methanation have been classified into two main ca-
tegories. The first one involves the conversion of CO2 to CO prior to
methanation, and the subsequent hydrogenation follows the same me-
chanism as CO mechanation [10,13]. The other reaction mechanism
Fig. 6. H2-TPR of all the samples. involves the direct hydrogenation of CO2 to CH4 without forming CO as
intermediate during the reaction [10,13]. Based on the conversion-de-
catalysts was enhanced dramatically. A trace amount of CO was de- pendent CO selectivity result of RT001 catalysts obtained at different
tected on both Ru-loaded catalysts at a low temperature of 150 °C. The space velocities, CO was identified to be an important intermediate for
formation of CH4 was observed above 175 °C. The yield of CH4 the CO2 methanation, which is further considered to occur in a CO-

247
S. Chai, et al. Journal of CO₂ Utilization 33 (2019) 242–252

Fig. 7. (a) CO2 conversion of all the samples.


(b) CH4 selectivity of all the samples. (c) CO
selectivity of all the samples. (d) CH4 formation
rate of all the samples. Reaction conditions:
pretreatment temperature 500 °C with 30 ml/
min H2 for 1 h, the feed flow rate amounted to
20 ml/min (10 vol.% CO2, 40 vol.% H2 diluted
in N2).

Table 2
Chemical parameters for the catalysts.
Sample Ru (wt%)a Ru dispersion (%)b CH4 formation rate (μmol·s−1·g−1)c TOF (10−2·s−1)d Ea (kJ·mol−1)e

RT001 2.39 32.0 5.12 7.24 69.8


RT101 2.41 51.1 3.39 2.98 74.6

a
Values determined by ICP.
b
Values calculated based on the H2-TPD results.
c
The reaction rate(r) was defined as the mole number of CH4 produced per gram of catalyst per second at 300 °C.
d
the TOF values was calculated at 300 °C.
e
Values calculated from the Arrhenius plots.

Fig. 8. Arrhenius plots of CO2 hydrogenation


over the RT001 and RT101 catalysts. Reaction
conditions: 200 mg catalyst were tested with a
feed flow rate of 20 ml/min (10 vol.% CO2,
40 vol.% H2 diluted in N2) between 150 and
190 °C. (b) Long term test of the RT001 sample
for CO2 conversion and CH4selectivity.
Reaction conditions: 200 mg catalysts were
tested with a feed flow rate of 20 ml/min
(10 vol.% CO2, 40 vol.% H2 diluted in N2) at a
reaction temperature of 325 °C.

mediated cascade CO2→CO→CH4 reaction mechanism. This result undetectable at temperatures exceeding 200 °C. To further explore the
seems to be reasonable since the CO molecule is prone to be activated role of CO in this reaction, the effects of different feed composition and
comparing to inert molecule CO2. space velocity on the CH4 production rate and product distribution
were studied for a wide range of weight hourly space velocity values at
250 °C over the RT001 catalyst. Additional investigations were per-
3.3. The role of CO in CO2 methanation formed using CO as initial reactant. As shown in the Fig. 10, the CH4
formation rate varied in the following order: RT001 (CO+H2) > RT001
A considerable amount of CO had been detected at low reaction (CO2+H2) > RT101 (CO2+H2) > T001 (CO2+H2) ≅ T101 (CO2+H2).
temperatures in the range between 150 °C and 175 °C, but it became

248
S. Chai, et al. Journal of CO₂ Utilization 33 (2019) 242–252

study, the TOF of CH4 formation over RT001 was almost 2.5 times
higher than the value over RT101 (see Table 2). Our reactivity data also
indicated that a larger Ru cluster size is more favorable for CH4 for-
mation, agreeing well with literature reports on other Ru-based cata-
lysts for CO2 methanation [22,27]. Based on our activity data and
characterization results, we concluded that the nature of the TiO2 cat-
alyst support has a strong impact on the dispersion of the Ru species.
To shed further light on the origin of the superior activity of the
RT001 catalyst, the chemisorption behaviors of CO2 and H2 over all the
samples were investigated in more detail by CO2-TPD and CO2-TPSR,
respectively. Fig. 11 shows the CO2 desorption profiles obtained from
CO2-TPD over RT001 and RT101. The results of CO2-TPD showed the
significantly stronger CO2 adsorption over RT001 than those on RT101.
The strong peak of CO2 desorption could be observed at 84 °C and 93 °C
well below 150 °C over RT001 and RT101, respectively. By contrary, no
obvious desorption peaks of CO2 and CO were detected at the initial
temperature of the CO2-TPD profiles on the bare supports of T001 and
T101 (Fig. S4), while only water was desorbed from the surface of the
Fig. 9. The effect of space velocity (L·h−1·g−1) on the conversion and product two supporting materials, in line with the activity data. The peak in-
selectivity of CO2 methanation at 250 °C.Reaction conditions: catalysts were tensity of adsorption CO2 over RT001 is much more intense than that on
tested with a feed flow rate of 20 ml/min (10 vol.% CO2, 40 vol.% H2 diluted in the pure support, indicating that CO2 is effectively adsorbed over the
N2). sites located at the interface between Ru and TiO2 in RT001 catalysts
[69], and the strong interaction of metal and support promotes CO2
adsorption. Previous studies [70,71] have demonstrated that the CO2
methanation requires the cooperation of metal site which is able to
dissociate H2, and the oxygen vacancy located in metal-support inter-
face which is able to activate CO2. It has been well documented in the
literatures that the high oxygen vacancy provides active sites to benefit
CO2 adsorption and activation and further improve the catalytic ac-
tivity for CO2 conversion [72–76]. For instance, a combined theoretical
and experimental result by Chen et al. show that CO2 binding is
strengthened by 0.32 eV compared to Pt when an oxygen vacancy is
generated at Pt–TiO2 interface, indicating that the activation of CO2
occurs at metal-support interface with oxygen vacancy generating by
the metal-support interaction and the enhanced activity of Pt/oxide for
CO2 reduction is originated from the sites at the Pt–oxide interface,
where the synergy between Pt and oxide plays an important role [76].
Fig. 12 displays the CO2-TPSR profiles of the RT001 and RT101
catalysts under the atmosphere of 10% H2/Ar after adsorption of CO2. It
shows again that the desorption peak of CO2 on the RT001 catalysts is
more intensive than on the RT101 catalysts, which is in accordance
Fig. 10. Comparison of the reaction rate of CO2 methanation over the RT001
with the results of CO2-TPD. Moreover, for the signals of CH4 were
and RT101 samples and of the reaction rate of CO methanation over the RT001
shown in Fig. 12 (b), we can see that the peak intensity of CH4 on the
sample. Reaction conditions: 200 mg catalysts were tested with a feed flow rate
of 20 ml/min (10 vol.% CO2, 40 vol.% H2 diluted in N2). RT001 catalyst is obviously also stronger than over the RT101 catalyst
during the whole desorption process, demonstrating that RT001 cata-
lyst is able to more efficiently promote CH4 formation via effective CO2
It is obvious that CO methanation is more readily to take place com- activation. No desorption of CO was observed as it gets visible in
pared with CO2 methanation, implying that the CO hydrogenation step Fig. 12(c). H2O is another product of CO2 methanation also during the
is not the rate-limiting step for cascade CO2 conversion. TPD experiments (see Fig. 12(d)). The release of H2O takes place in the
same temperature range as for CO2 which further confirms the occur-
3.4. The impact of underlying TiO2 support on CO2 methanation rence of CO2 methanation over the Ru/TiO2 samples. The CO2-TPSR
profile of T001 and T101 are displayed in Fig. S6. No obvious deso-
All the relevant chemical parameters for the RT001 and RT101 rption peaks of CO2, CH4 or CO could be observed during the TPD
catalysts are listed in Table 2. The actual Ru loading is determined to be process and only some desorbed from the surface over the two supports
2.39% over the T001 and 2.41% over the T101 supports, respectively, similar to the result of CO2-TPD, illustrating the low reactivity of T001
which are both very close to the nominal loading of added Ru pre- and T101 for the CH4 formation despite the high capability of CO2
cursors. Only a very small amount of ruthenium is lost during the cat- adsorption on the T001 surface. In this CO2-TPSR experiment, the Ru-
alyst preparation. The value of Ru dispersion over the T101 support is containing catalysts exhibited a remarkable ability for CH4 formation.
remarkably higher than over the T001 support, 51.11% versus 32.03% Oxygen vacancy has long been considered as the reactive site on the
respectively, indicating that the size of the Ru nanoparticles is much surface of metal oxides [10]. In Ru/TiO2 catalysts, the oxygen vacancy
smaller for the T101 sample than that of the T001 sample. The che- may act as an important site in adsorbing and activating carbon-oxygen
misorption results were also consistent with the results obtained from bond. Previous reports have revealed the close relationship between the
TEM measurements. Recently, Kwak et al. have reported that the ac- concentration of oxygen vacancy and the metal–support interaction,
tivity and selectivity of Ru/Al2O3 catalysts were very sensitive to the illustrating the metal-support interaction at metal–support interface
cluster size of the Ru nanoparticles dispersed on the support, large Ru plays a key role in the concentration of oxygen vacancy [20,77]. The
clusters tended to favor the CH4 formation [22,27]. In the present loading of Ru onto CeO2 has been found to greatly promote the

249
S. Chai, et al. Journal of CO₂ Utilization 33 (2019) 242–252

Fig. 11. Mass signals of (a) CO2 (m/e = 44), (b) CO (m/e = 28), (c) H2O (m/e = 18) and (d) CH4 (m/e = 15) during the CO2-TPD process over RT001 and RT101
after CO2 preadsorption.

Fig. 12. Mass signals of (a) CO2 (m/e = 44), (b) CH4 (m/e = 15), (c) CO (m/e = 28), and (d) H2O (m/e = 18) during the CO2-TPSR process over RT001 and RT101.
Determined under H2 atmosphere during the temperature ramp after CO2 preadsorption.

250
S. Chai, et al. Journal of CO₂ Utilization 33 (2019) 242–252

formation of oxygen vacancy. Wang et al. reported that the surface facet of support in Ru/TiO2 catalysts were identified to be one of the
oxygen vacancy rather than Ru metal in Ru/CeO2 catalyst is the active effective strategies to enhance the population of surface oxygen va-
site for the rate-controlling step of CO2 methanation and abundant cancies in facilitating CO2 adsorption and thereby boost CO2 metha-
oxygen vacancies at the interface of Ru and CeO2-NCs could facilitate nation efficiencies. The highly enhanced catalytic activity of Ru species
the CO2 activation [20]. Our results of Raman, CO2-TPD, and CO2-TPSR loaded on TiO2 (001) facets was originated from the nature of (001)
presented in this work suggest that the strong interaction of Ru and facets with high surface energy and abundant oxygen vacancies, which
TiO2 (001) lead to highly enhanced concentration of oxygen vacancies could provide more active sites for CO2 activation and result in the
located in the interface of Ru and support, and facilitate the adsorption synergistic effect of ruthenium and TiO2 due to the strong metal-sup-
of CO2 as well as the activation of C]O inert bond, thus promoting the port interaction. Therefore, RT001 catalysts displayed an excellent ac-
overall reactivity owing to the facet-dependent and Ru-promoted for- tivity and durability, demonstrating its promising potential for appli-
mation of oxygen vacancy. cations in hydrogenation of carbon dioxide.
In terms of reaction mechanism, our conversion-dependent CO se-
lectivity result indicates a CO-mediated cascade CO2→CO→CH4 reac- Acknowledgements
tion mechanism over Ru supported TiO2 (001). According to the test of
CO methanation over RT001, the activation of CO2, rather than CO This research was supported by Science and Technology
hydrogenation step, seems to be a more kinetic-relevant step for cas- Commission of Shanghai Municipality (18030501100), the Program for
cade CO2→CO→CH4 transformation. This is also consistent with the Professor of Special Appointment (Eastern Scholar) at Shanghai
results of the morphology/facet-dependent and Ru-promoted oxygen Institutions of Higher Learning, the Talent Program of Shanghai
vacancy obtained from CO2-TPD and Raman measurements. University of Engineering Science, National Natural Science Foundation
Early studies found the loading of small Ru metal nanoparticles on of China (21503133), and Shanghai Talent Development Foundation
the reducible supports (e.g., TiO2, CeO2) can significantly enhance their (2017076).
COx methanation performance owing to the metal–support interactions
[10,77]. The activity and selectivity of these catalysts have been shown Appendix A. Supplementary data
to be very sensitive to the cluster size and the interaction between the
active metals and oxide supports. In particular, for Ru/TiO2 catalysts, Supplementary material related to this article can be found, in the
TOFs of both CO and CO2 methanation increase by a factor of 40 and online version, at doi:https://doi.org/10.1016/j.jcou.2019.05.031.
25, respectively, with increasing mean particle size of Ru from 2.1 to
4.5 nm, accompanied by an increase of methane selectivity [78]. The References
observed dependence of TOF on crystallite size indicates that hydro-
genation of CO/CO2 occurs preferably on flat Ru surfaces, the fraction [1] J. Hansen, M. Sato, R. Ruedy, K. Lo, D.W. Lea, M. Medina-Elizade, Proc. Natl. Acad.
of which is higher for larger metal crystallites. A late study by Kwak Sci. 103 (2006) 14288–14293.
[2] N. MacDowell, N. Florin, A. Buchard, J. Hallett, A. Galindo, G. Jackson,
et al. dealing with size effect of Ru over Al2O3 reported the dramatic C.S. Adjiman, C.K. Williams, N. Shah, P. Fennell, Energ. Environ. Sci. 3 (2010)
decrease in CO and increase in CH4 selectivity with increasing Ru 1645–1669.
particle size, confirming the size-dependent activity and selectivity [3] T.R. Knutson, R.E. Tuleya, J. Climate 17 (2004) 3477–3495.
[4] D. Jo, J.B. Lim, T. Ryu, I.S. Nam, M.A. Camblor, S.B. Hong, J. Mater. Chem. A 3
[22]. (2015) 19322–19329.
The results that CO2 methanation is favored over Ru loaded on TiO2 [5] R.N.E. Huaman, T.X. Jun, Renew. Sustain. Energ. Rev. 31 (2014) 368–385.
(001) facets in this study may be explained by the decreased fraction of [6] K.M. Yu, I. Curcic, J. Gabriel, S.C. Tsang, ChemSusChem 1 (2008) 893–899.
[7] J. Rossmeisl, Z.W. Qu, H. Zhu, G.J. Kroes, J.K. Nørskov, J. Electroanal. Chem. 607
coordinately unsaturated Ru atoms at edges and corners relative to the (2007) 83–89.
total amount of surface metal atoms with increasing Ru crystallite size, [8] B. Girginer, G. Galli, E. Chiellini, N. Bicak, Int. J. Hydrogen Energ. 34 (2009)
which can facilitate the formation of CH4 by supplying large amounts of 1176–1184.
[9] C.S. Budi, H.C. Wu, C.S. Chen, D. Saikia, H.M. Kao, ChemSusChem 9 (2016)
atomic hydrogen to the hydrogenation step. Moreover, the Ru with
2326–2331.
large particle size and low unsaturation in the sample of RT001 pro- [10] W. Wang, S.P. Wang, X.B. Ma, J.L. Gong, Chem. Soc. Rev. 40 (2011) 3703–3727.
mote the formation of abundant surface oxygen vacancies, in compar- [11] J. Ma, N.N. Sun, X.L. Zhang, N. Zhao, F.K. Xiao, W. Wei, Y.H. Sun, Catal. Today 148
ison with Ru loaded over TiO2 (101) facets. The observed highly en- (2009) 221–231.
[12] T. Le, A. Striolo, C.H. Turner, D.R. Cole, Sci. Rep. 7 (2017) 9021.
hanced methanation activity in the catalyst of Ru loaded on TiO2 (001) [13] M.A.A. Aziz, A.A. Jalil, S. Triwahyono, A. Ahmad, Green Chem. 17 (2015)
is due, most probably, to the synergistic effect of concerted reaction 2647–2663.
between the oxide-activated CO2 and the metal-activated hydrogen, [14] H.C. Wu, Y.C. Chang, J.H. Wu, J.H. Lin, I.K. Lin, C.S. Chen, Catal. Sci. Technol. 5
(2015) 4154–4163.
owing to the strong metal-support interaction. Based on previous re- [15] Y. Yan, Y.H. Dai, H. He, Y.B. Yu, Y.H. Yang, Appl. Catal. B 196 (2016) 108–116.
ports and the results in this work, it is concluded that oxygen vacancy [16] J.K. Kesavan, I. Luisetto, S. Tuti, C. Meneghini, G. Iucci, C. Battocchio, S. Mobilio,
located in the Ru/TiO2 interface serves as the active site for activation S. Casciardi, R. Sisto, J. CO2. Util. 23 (2018) 200–211.
[17] Y. Yan, Y.H. Dai, Y.H. Yang, A.A. Lapkin, Appl. Catal. B 237 (2018) 504–512.
and conversion of CO2 to CO, while Ru provides site for CH4 production [18] D. Wierzbicki, R. Baran, R. Dębek, M. Motak, M.E. Gálvez, T. Grzybek, P. Da Costa,
via CO hydrogenation and regeneration of oxygen vacancy. P. Glatzel, Appl. Catal. B 232 (2018) 409–419.
[19] B. Mutz, H.W.P. Carvalho, S. Mangold, W. Kleist, J.D. Grunwaldt, J. Catal. 327
(2015) 48–53.
4. Conclusions [20] F. Wang, S. He, H. Chen, B. Wang, L.R. Zheng, M. Wei, D.G. Evans, X. Duan, J.Am.
Chem. Soc. 138 (2016) 6298–6305.
In this work, Ru nanoparticles supported by TiO2 (001) exhibited a [21] Y. Zhu, S.R. Zhang, Y.C. Ye, X.Q. Zhang, L. Wang, W. Zhu, F. Cheng, F. Tao, ACS
Catal. 2 (2012) 2403–2408.
highly enhanced CO2 conversion compared with TiO2 (101), and the
[22] J.H. Kwak, L. Kovarik, J. Szanyi, ACS Catal. 3 (2013) 2449–2455.
superior reactivity data by using TiO2 (001) surface can be satisfactorily [23] C.S. Kellner, A.T. Bell, J. Catal. 75 (1982) 251–261.
explained by the measurements based on HR-TEM, Raman spectra, H2- [24] G. Garbarino, D. Bellotti, E. Finocchio, L. Magistri, G. Busca, Catal. Today 277
TPR, and chemisorption measurements. Our activity data and char- (2016) 21–28.
[25] Y. Guo, S. Mei, K. Yuan, D.J. Wang, H.C. Liu, C.H. Yan, Y.W. Zhang, ACS Catal. 8
acterization results in this work provide evidences that CO2 methana- (2018) 6203–6215.
tion activity over Ru-based catalysts can be determined by the choice of [26] J.N. Park, E.W. McFarland, J. Catal. 266 (2009) 92–97.
support material and exposed crystal facets, which can affect not only [27] J.H. Kwak, L. Kovarik, J. Szanyi, ACS Catal. 3 (2013) 2094–2100.
[28] S.K. Beaumont, S. Alayoglu, C. Specht, N. Kruse, G.A. Somorjai, Nano Lett. 14
the size of supported Ru nanoparticles, but also the number of available (2014) 4792–4796.
surface oxygen vacancy effective for CO2 activation. As demonstrated [29] H.H. Shin, L. Lu, Z. Yang, C.J. Kiely, S. McIntosh, ACS Catal. 6 (2016) 2811–2818.
by activity and characterization results, modulating the exposed crystal [30] K.P. Yu, W.Y. Yu, M.C. Kuo, Y.C. Liou, S.H. Chien, Appl. Catal. B 84 (2008)

251
S. Chai, et al. Journal of CO₂ Utilization 33 (2019) 242–252

112–118. [54] J.H. Xu, X. Su, H.M. Duan, B.L. Hou, Q.Q. Lin, X.Y. Liu, X.L. Pan, G.X. Pei,
[31] A. Karelovic, P. Ruiz, Appl. Catal. B 113-114 (2012) 237–249. H.R. Geng, Y.Q. Huang, T. Zhang, J. Catal. 333 (2016) 227–237.
[32] J.C. Matsubu, V.N. Yang, P. Christopher, J. Am. Chem. Soc. 137 (2015) 3076–3084. [55] J.G. Wang, P.H. Rao, W. An, J.L. Xu, Y. Men, Appl. Catal. B 195 (2016) 141–148.
[33] T. Abe, M. Tanizawa, K. Watanabe, A. Taguchi, Energy Environ. Sci. 2 (2009) [56] D.P. Debecker, B. Farin, E.M. Gaigneaux, C. Sanchez, C. Sassoye, Appl. Catal. A Gen.
315–321. 481 (2014) 11–18.
[34] C. Sassoye, G. Muller, D.P. Debecker, A. Karelovic, S. Cassaignon, C. Pizarro, [57] J. Zhu, S.H. Wang, Z.F. Bian, S.H. Xie, C.L. Cai, J.G. Wang, H.G. Yang, H.X. Li,
P. Ruiz, C. Sanchez, Green Chem. 13 (2011) 3230–3237. CrystEngComm 12 (2010) 2219–2224.
[35] G.L. Bezemer, J.H. Bitter, H.P.C.E. Kuipers, H. Oosterbeek, J.E. Holewijn, X.D. Xu, [58] Z. Wang, C.Y. Yang, T.Q. Lin, H. Yin, P. Chen, D.Y. Wan, F.F. Xu, F.Q. Huang,
F. Kapteijn, A.J. van Dillen, K.P. de Jong, J. Am. Chem. Soc. 128 (2006) 3956–3964. J.H. Lin, X.M. Xie, M.H. Jiang, Adv. Funct. Mater. 23 (2013) 5444–5450.
[36] J.M. González-Carballo, F.J. Pérez-Alonso, M. Ojeda, F.J. García-García, [59] X.X. Huang, Y. Men, J.G. Wang, W. An, Y. Wang, Catal. Sci. Technol. 7 (2017)
J.L.G. Fierro, S. Rojas, ChemCatChem 6 (2014) 2084–2094. 168–180.
[37] A. Karelovic, P. Ruiz, J. Catal. 301 (2013) 141–153. [60] J.C. Parker, R.W. Siegel, J. Mater. Res. 5 (1990) 1246–1252.
[38] M. Behrens, F. Studt, I. Kasatkin, S. Kuhl, M. Havecker, F. Abild-Pedersen, [61] J.C. Parker, R.W. Siegel, Appl. Phys. Lett. 57 (1990) 943–945.
S. Zander, F. Girgsdies, P. Kurr, B.L. Kniep, M. Tovar, R.W. Fischer, J.K. Nørskov, [62] J. Zhu, J. Ren, Y.N. Huo, Z.F. Bian, H.X. Li, J. Phys. Chem. C. 111 (2007)
R. Schlögl, Science. 336 (2012) 893–897. 18965–18969.
[39] X. Liu, Y. Men, J.G. Wang, R. He, Y.Q. Wang, J. Power Sources 364 (2017) 341–350. [63] S.J. Tauster, S.C. Fung, R.L. Garten, J. Am. Chem. Soc. 100 (1978) 170–175.
[40] A. Kim, D.P. Debecker, F. Devred, V. Dubois, C. Sanchez, C. Sassoye, Appl. Catal. B [64] S.J. Tauster, Acc. Chem. Res. 20 (1987) 389–394.
220 (2018) 615–625. [65] G.L. Haller, D.E. Resasco, Adv. Catal. 36 (1989) 173–235.
[41] W.H. Li, X.W. Nie, X. Jiang, A.F. Zhang, F.S. Ding, M. Liu, Z.M. Liu, X.W. Guo, [66] Z. Mohamed, V.D. Dasireddy, S. Singh, H.B. Friedrich, Int. J. Hydrogen Energ. 43
C.S. Song, Appl. Catal. B 220 (2018) 397–408. (2018) 22291–22302.
[42] A. Kim, C. Sanchez, G. Patriarche, O. Ersen, S. Moldovan, A. Wisnet, C. Sassoye, [67] D.I. Kondarides, P. Panagiotopoulou, X.E. Verykios, J. Phys. Chem. C 115 (2011)
D.P. Debecker, Catal. Sci. Technol. 6 (2016) 8117–8128. 1220–1230.
[43] Y. Li, W.J. Shen, Chem. Soc. Rev. 43 (2014) 1543–1574. [68] S. Bernal, F.J. Botana, J.J. Calvino, C. López, J.A. Pérez-Omil, J.M. Rodríguez-
[44] G. Liu, H.G. Yang, J. Pan, Y.Q. Yang, G.Q. Lu, H.M. Cheng, Chem. Rev. 114 (2014) Izquierdo, J. Am. Chem. Soc. 92 (1996) 2799–2809.
9559–9612. [69] A. Beuls, C. Swalus, M. Jacquemin, G. Heyen, A. Karelovic, P. Ruiz, Appl. Catal. B
[45] L. Cheng, Y. Men, J.G. Wang, H. Wang, W. An, Y.Q. Wang, Z.C. Duan, J. Liu, Appl. 113–114 (2012) 2–10.
Catal. B 204 (2017) 374–384. [70] J.A.H. Dreyer, P. Li, L. Zhang, G.K. Beh, R. Zhang, P.H.L. Sit, W.Y. Teoh, Appl.
[46] F. Ji, Y. Men, J.G. Wang, Y.L. Sun, Z.D. Wang, B. Zhao, X.T. Tao, G.J. Xu, Appl. Catal. B 219 (2017) 715–726.
Catal. B 242 (2019) 227–237. [71] D. Li, N. Ichikuni, S. Shimazu, T. Uematsu, Appl. Catal. A Gen. 180 (1999) 227–235.
[47] A. Vittadini, A. Selloni, F.P. Rotzinger, M. Grätzel, Phys. Rev. Lett. 81 (1998) [72] J.Y. Ye, C.J. Liu, D.H. Mei, Q.F. Ge, ACS Catal. 3 (2013) 1296–1306.
2954–2957. [73] Y.X. Pan, C.J. Liu, D.H. Mei, Q.F. Ge, Langmuir 26 (2010) 5551–5558.
[48] J.G. Wang, Z.F. Bian, J. Zhu, H.X. Li, J. Mater. Chem. A 1 (2013) 1296–1302. [74] S.C. Yang, W.N. Su, J. Rick, S.D. Lin, J.Y. Liu, C.J. Pan, J.F. Lee, B.J. Hwang,
[49] X.Q. Gong, A. Selloni, A. Vittadini, J. Phys. Chem. B 110 (2006) 2804–2811. ChemSusChem 6 (2013) 1326–1329.
[50] L. Mino, G. Spoto, A.M. Ferrari, J. Phys. Chem. C. 118 (2014) 25016–25026. [75] S. Kattel, P. Liu, J.G. Chen, J. Am. Chem. Soc. 139 (2017) 9739–9754.
[51] Y.H. Cao, Q.Y. Li, C. Li, J.L. Li, J.J. Yang, Appl. Catal. B 198 (2016) 378–388. [76] S. Kattel, B.H. Yan, J.G. Chen, P. Liu, J. Catal. 343 (2016) 115–126.
[52] J. Mao, L.Q. Ye, K. Li, X.H. Zhang, J.Y. Liu, T.Y. Peng, L. Zan, Appl. Catal. B 144 [77] F. Wang, C.M. Li, X.Y. Zhang, M. Wei, D.G. Evans, X. Duan, J. Catal. 329 (2015)
(2014) 855–862. 177–186.
[53] Q.Q. Lin, X.Y. Liu, Y. Jiang, Y. Wang, Y.Q. Huang, T. Zhang, Catal. Sci. Technol. 4 [78] P. Panagiotopoulou, D.I. Kondarides, X.E. Verykios, Appl. Catal. B 88 (2009)
(2014) 2058–2063. 470–478.

252

You might also like