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

pubs.acs.

org/JPCC Article

Selective Hydrogenation of CO2 to Ethanol over Sodium-Modified


Rhodium Nanoparticles Embedded in Zeolite Silicalite‑1
Published as part of The Journal of Physical Chemistry virtual special issue “Energy and Catalysis in China”.
Fuyong Zhang,† Wei Zhou,† Xuewei Xiong, Yuhao Wang, Kang Cheng, Jincan Kang,* Qinghong Zhang,
and Ye Wang*
Cite This: J. Phys. Chem. C 2021, 125, 24429−24439 Read Online
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via INDIAN INST OF TECH DELHI- IIT on December 27, 2021 at 07:36:52 (UTC).

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: Catalytic transformation of CO2 into chemicals in large


demand such as ethanol has attracted much research attention under the
background of establishing carbon-neutral societies. Supported Rh
catalysts are promising candidates for the hydrogenation of CO2 to
ethanol but suffer from low ethanol productivity and poor catalyst stability.
Here, we report that zeolite silicalite-1 embedded Na-promoted Rh
nanoparticles (Na-Rh@S-1) demonstrate high productivity and stability
for CO2 hydrogenation to ethanol. The ethanol selectivity of 24% was
attained at a CO2 conversion of 10%, and the space-time yield of ethanol
reached 72 mmol gRh−1 h−1, which outperformed most of the Rh-based
catalysts reported to date. While a reference catalyst prepared by
impregnation underwent deactivation, the Na-Rh@S-1 catalyst was stable
for at least for 100 h owing to the confinement effect. The Na+ modifier played crucial roles in enhancing the CO2 conversion and
ethanol selectivity by suppressing methane formation. The characterizations suggest that the presence of Na+ enables the coexistence
of Rh0 and Rh+ and enhances CO2 adsorption, thus boosting ethanol formation. A comparative study between CO and CO2
hydrogenation reveals that the Na-Rh@S-1 catalyst is significantly more active and selective toward CO2 hydrogenation to ethanol.

1. INTRODUCTION ethanol has attracted much research attention, but the


The capture and recycling of CO2 released during industrial productivity of ethanol remains low.16−19 This is not only
processes to chemical and energy feedstocks in great demand because CO2 is a thermodynamically stable and kinetically
would contribute to establishing carbon-neutral societies, and inert molecule but also because the formation of C2H5OH is
thus have become a hot research area.1,2 The hydrogenation of generally believed to involve the C−C coupling between
CO2 by “green” H2, which can be produced by electrolysis of different species (e.g., CHx and CHxO (x = 0−3)), which is
H2O using renewable electricity,3 into high-value chemicals is difficult to control.16,19 Among the reported catalysts for CO2
one of the most promising routes for the utilization of CO2.4−6 hydrogenation to C2+ oxygenates, mainly including Rh-,20−24
Methanol has long been an important target product for CO2 Cu-,25,26 Co-,27,28 and Mo-based catalysts,29 the Rh-based
hydrogenation,7,8 but the direct transformation of CO2 catalysts show promising performance. It is well-known that
selectively into a more value-added multicarbon (C2+) Rh-based catalysts are also among the most efficient catalysts
compound is also very attractive and is more challenging for the conversion of syngas (CO + H2) to ethanol owing to
because the control of C−C coupling is a big challenge in C1 the unique abilities of Rh in C−O bond activation and CO
chemistry.6,9 Significant progress has been achieved for the insertion for C−C coupling.30,31 As compared to Rh-catalyzed
hydrogenation of CO2 into C2+ hydrocarbons, including lower CO hydrogenation, the studies on CO2 hydrogenation over
olefins,10,11 aromatics,11,12 and liquid hydrocarbon fuels,13 by Rh-based catalysts are still limited. Similar to CO hydro-
developing new strategies such as tandem or relay catalysis. genation, the CH4 or CO dominated the product in CO2
However, the selective synthesis of C2+ oxygenates by CO2
hydrogenation still remains a big challenge. Received: September 5, 2021
As one of the most demanded C2+ oxygenates, ethanol is Revised: October 10, 2021
considered as an ideal fuel additive, alternative fuel, attractive Published: October 24, 2021
hydrogen carrier, and versatile building-block chemical for the
synthesis of value-added chemicals, advanced fuels, and
polymers.14,15 Recently, the hydrogenation of CO2 into

© 2021 American Chemical Society https://doi.org/10.1021/acs.jpcc.1c07862


24429 J. Phys. Chem. C 2021, 125, 24429−24439
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

hydrogenation over unmodified Rh catalysts. Promoters, such (EDTA-2Na), and sodium nitrate (NaNO3, 99%) were
as the alkali metal ions (e.g., Li+, Na+, and K+) and transition purchased from Sinopharm Chemical Reagent Co. Rhodium
metal oxides (e.g., FeOx, MnOx, and VOx), and supports chloride hydrate (RhCl3·xH2O, Rh content, 38.5−42.5%) was
played key roles in tuning the electronic state of Rh species and purchased from Aladdin.
the metal−support interactions to boost the formation of The nanosized S-1 zeolite was synthesized by a hydro-
ethanol.16,19−24 However, the state-of-the-art ethanol produc- thermal method. The molar ratio of the components in the
tivity for CO2 hydrogenation is still far from satisfactory. synthesis gel was 1 SiO2: 0.4 TPAOH: 35 H2O. For typically
Moreover, the supported Rh catalysts usually display poor procedures, 13.0 g of TPAOH was first dissolved in 15.42 g of
stability due to the sintering of nanoparticle during CO2 deionized water under vigorous stirring, and then 8.32 g of
hydrogenation.16 Therefore, new strategies should be devel- TEOS was added dropwise into the above solution. After the
oped to increase both the productivity and stability of Rh- TEOS was hydrolyzed completely, the mixed gel was
based catalysts for CO2 hydrogenation to ethanol. Further- transferred into a 100 mL Teflon-lined autoclave for
more, it is of fundamental interest to understand the hydrothermal crystallization. The hydrothermal synthesis was
similarities and differences between Rh-catalyzed CO2 and carried out at 443 K for 96 h. The resultant solid was collected
CO hydrogenation reactions, but such comparative studies are by centrifugation, thoroughly washed with water 5 times, and
very rare.32 subsequently washed with ethanol. The solid product was
Zeolites, which possess well-defined crystalline microporous dried at 353 K overnight and calcined in air at 823 K for 8 h.
structures and high thermal/hydrothermal stability, have The Rh@S-1 catalyst, which contained Rh nanoparticles
received renewed interests in catalysis because of recent fixed in nanosized S-1 zeolite, was synthesized by a modified
successful synthesis and the unique catalytic behaviors of method reported previously.37 EDA was employed as the
metal@zeolite materials.33−36 The active metal species (either protect ligand. The procedures were similar to those adopted
single-atom and few-atom clusters smaller than the micropores for the synthesis of S-1 except for the addition of an aqueous
or nanoparticles larger than the micropores) in the metal@ solution of [Rh(NH2CH2CH2NH2)3]Cl3, which was prelimi-
zeolite materials are needlessly located inside zeolite micro- narily prepared by dissolving 0.063 g of RhCl3·3H2O and 0.8
pores but can be embedded in situ in the interrupted zeolite mL of EDA in 5 mL of deionized H2O, into the synthesis gel.
frameworks during their formation and are fixed by the zeolite The Na-Rh@S-1 catalyst (i.e., the Na-modified Rh nano-
frameworks. This offers more opportunities for designing particles fixed in nanosized S-1) was also synthesized by similar
different types of metal@zeolite hybrid catalysts with unique procedures to those of S-1 or Rh@S-1. EDTA was chosen as
catalytic behaviors such as shape selectivity and high the ligand of Rh in the presence of Na ions. A mixed solution
stability.33−36 For example, a ligand-protected hydrothermal containing Rh-EDTA complex and NaCl was preliminarily
crystallization was developed to synthesize subnanometer Pt− prepared by dissolving 0.063 g of RhCl3·3H2O and 0.10 g of
Zn clusters confined in silicalite-1 (S-1), a pure silicious zeolite EDTA-2Na into 5 mL of deionized H2O, and this solution was
with MFI structure, and the Pt−Zn@S-1 displayed high added into the synthesis gel. The content of Na in the final
activity and stability for the dehydrogenation of propane to catalyst was tuned by changing the repeated time used for
propylene at 823 K.37 A Pd@S-1 catalyst with an average Pd water-washing step. The samples with different contents of Na
particle size of 2.3 nm synthesized by the same ligand- were denoted as xNa-Rh@S-1, where x was the weight content
protected hydrothermal method showed high activity and of Na.
stability in methane combustion reaction.38 Rh−Mn nano- A reference catalyst (i.e., the S-1-supported catalyst
particles were confined within S-1 by a metal-containing seed- (denoted as Na-Rh/S-1)) was prepared by the conventional
directed crystallization, and the RhMn@S-1 catalyst showed incipient wetness impregnation method. Typically, 0.004 g of
high activity and stability in the conversion of syngas to C2+ NaNO3 and 0.012 g of RhCl3·3H2O were dissolved in 3.0 g of
oxygenates.39 Despite being applied successfully in several deionized water under vigorous stirring for 30 min to form a
types of reactions, the studies on the application of metal@ clear solution, and then 0.60 g of S-1 was slowly added into the
zeolite catalysts for CO2 hydrogenation to C2+ oxygenates are mixed aqueous solution under stirring for 10 min. The
rare.33−36 obtained slurry was further dried at 353 K for 10 h, followed by
Here, we report our recent studies on the application of Na- calcination in air at 773 K for 5 h to obtain the catalyst
modified Rh nanoparticles embedded in S-1 (denoted as Na- denoted as 0.17Na-Rh/S-1.
Rh@S-1) synthesized by the ligand-protected hydrothermal 2.2. Catalyst Characterization. X-ray diffraction (XRD)
crystallization in CO2 hydrogenation to ethanol. We have also patterns were recorded on a Rigaku Ultima IV diffractometer
carried out a comparative study between hydrogenation of CO (Rigaku, Japan) using Cu Kα radiation (40 kV and 30 mA) as
and CO2 over the S-1 confined Rh catalyst. The role of Na+ in the X-ray source. N2 physisorption measurements were carried
the catalyst and the effect of confinement on catalytic out on a Micromeritics ASAP 2460 and the sample was
performances including the stability have also been inves- pretreated under vacuum at 473 K for 2 h prior to each
tigated to gain insights into nature of active sites for selective measurement. The contents of Rh and Na in the catalyst were
formation of ethanol and the advantage of the metal@zeolite measured by inductively coupled plasma-optical emission
catalytic materials. spectrometer (ICP-OES). X-ray photoelectron spectroscopy
(XPS) measurements were carried out in an UHV chamber
2. METHODS equipped with an Omicron XPS (base pressure 5 × 10−10
2.1. Materials and Catalyst Synthesis. All the reagents Torr). A monochromatized Al Kα X-ray source and a Sphere 2
were purchased and used without further purification. analyzer were used. After pretreatment in H2 (5 vol %)/Ar flow
Tetraethyl orthosilicate (TEOS), tetrapropylammonium hy- at 673 K for 30 min in a pretreatment chamber directly
droxide aqueous solution (TPAOH, 25 wt %), ethylenedi- connected to the detecting chamber, the sample was
amine (EDA), ethylenediaminetetraacetic acid disodium salt transferred into the UHV detecting chamber for XPS
24430 https://doi.org/10.1021/acs.jpcc.1c07862
J. Phys. Chem. C 2021, 125, 24429−24439
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 1. Schematic diagram for the formation of Na-Rh@S-1 catalyst.

measurements. To obtain the surface information of the Rh- or gas flow (30 mL min−1) at 673 K for 30 min in the DRIFTS
Na-promoted Rh nanoparticles embedded in zeolite, the Ne cell. Then, the sample was cooled to 523 K, followed by
ions were used to sputter the sample with a kinetic energy of 5 purging with He. After recording the background spectrum, a
keV, an ion flux of 1600 pA·m−2, and a spot size of 2 mm × 2 CO2/H2 (24 vol % CO2/72 vol % H2/4 vol % Ar) gas at 2.5
mm, which could strip off the zeolite shell. Then, the sputtered MPa was switched into the in situ DRIFTS cell at 523 K, and
sample was transferred into the analysis chamber for XPS the evolution of IR spectra was recorded. The in situ DRIFTS
measurements to obtain the surface information on the measurements were carried out on a Bruker Vertex 70
embedded nanoparticles. instrument equipped with a high-pressure Harrick Scientific
Scanning electron microscopy (SEM) measurements were DRIFTS cell. After pretreatment with H2 at 673 K followed by
carried out with a Hitachi S-4800 operated at 15 kV. He purging and cooling to 523 K, high-pressure CO2 (2.5
Transmission electron microscopy (TEM), dark-field TEM, MPa) was first introduced to the cell, and then the gas feed was
energy dispersive spectroscopy (EDS) mapping, and line-scan switched to high-pressure H2 (2.5 MPa). The evolution of IR
elemental analysis measurements were carried out on a Phillips spectra was recorded at different times after switching the gas
Analytical FEI Tecnai 20 electron microscope operated at an feed to H2.
acceleration voltage of 200 kV. The ultramicrotomy of the 2.3. Catalytic Reaction. The CO2 and CO hydrogenation
resin-embedded catalyst was carried out to expose cross reactions were both carried out on a high-pressure, fixed-bed
sections of the catalyst with metal nanoparticles confined or titanium reactor with an inner diameter of 10 mm designed by
fixed in S-1. For this purpose, the sample was first embedded in Xiamen HanDe Engineering Co., Ltd. Typically, a catalyst
epoxy resin, and then cut into 60 nm sections using a diamond (0.15 g) with grain sizes of 30−60 mesh was loaded in the
knife. Slices of the sample were deposited on carbon-coated titanium reactor. Prior to each reaction, the catalyst was
copper grids (200 meshes) for TEM measurements.40,41 pretreated in H2(5 vol %)/Ar (25 mL min−1) at 673 K for 30
Hydrogen temperature-programmed reduction (H2-TPR) min. Then the reactant gases (24%CO2/72%H2/4%Ar) were
measurements were carried out on a Micromeritics AutoChem introduced into the titanium reactor with 15 mL min−1 at 523
II 2920 instrument. Prior to each measurement, 0.10 g of K and 5.0 MPa. Products were analyzed by an on-line gas
sample was pretreated in a He flow at 673 K for 1 h. Then, the chromatograph (Ruimin GC 2060, Shanghai) equipped with a
sample was cooled down to 323 K, followed by switching to H2 thermal conductivity detector (TCD) and a flame ionization
(5 vol %)/Ar gas flow. Subsequently, the temperature was detector (FID). A TDX-01-packed column was connected to
increased to 773 K at a ramping rate of 10 K min−1, and the H2 the TCD for the separation and analyses of Ar, CO, CH4, and
consumption was detected by a thermal conductivity detector CO2. A RT-Q-BOND-PLOT capillary column was connected
(TCD). The chemisorption of CO2 was measured volumetri- to the FID for the separation and analyses of organic
cally on a Micromeritics 3flex instrument. Prior to each oxygenated compounds (including dimethyl ether, methanol,
measurement, the 0.10 g sample was pretreated in a H2 (5%)/ ethanol, and propanol) and C1−C10 hydrocarbons. The
N2 flow at 673 K for 30 min, followed by degassing in vacuum product selectivity was calculated on a molar carbon basis.
at 423 K for 2 h. The chemisorption of CO2 was detected at The selectivity of CO formed via the reverse water−gas shift
298 K by a Rubotherm Isosorp-Htgra magnetic suspension (RWGS) reaction was calculated separately from the hydro-
balance. genated products. Carbon balances were all in the range of
CO-adsorbed Fourier-transform infrared (CO-FTIR) meas- 95−99%.
urements were carried out on a Nicolet 6700 instrument
equipped with a MCT detector. The spectra were recorded 3. RESULTS AND DISCUSSION
from 4000 to 600 cm−1 at a resolution of 2 cm−1. The sample 3.1. Morphology and Structure of Catalysts. Figure 1
was pressed into a self-supported wafer and was set in an in situ illustrates the overall process for the formation of Na-modified
IR cell. For a typical measurement, after the sample was Rh nanoparticles confined within S-1 zeolite. In brief, the
pretreated in a H2 (5 vol %)/Ar flow at 673 K for 30 min, the ligand EDTA forms a complex with Rh in the presence of Na+
sample was cooled down to room temperature, and then the and can avoid the precipitation of Rh species under strong
adsorption of CO was conducted for 30 min by introduction of alkaline conditions for hydrothermal crystallization of S-1.
a CO (10 vol %)/Ar flow to the cell. The CO-FTIR spectrum Furthermore, the Rh complex does not affect the strong
was collected after the removal of gaseous CO molecules by alkaline environment for the synthesis of S-1 in the presence of
evacuation. TPAOH (the template), and thus would not affect the
In situ high-pressure diffuse reflectance infrared Fourier- crystallization process of S-1. After hydrothermal synthesis and
transform spectroscopy (DRIFTS) measurements were carried the removal of organic template by calcination, the Rh species
out on the same instrument equipped with a high-pressure and Na+ ions would be fixed by the framework of S-1, thus
Harrick Scientific DRIFTS cell of ZnSe windows. Prior to each forming the Na+-modified Rh nanoparticles embedded within
measurement, the sample was pretreated in H2 (5 vol %)/Ar S-1 zeolite. The Na content in catalysts was regulated by
24431 https://doi.org/10.1021/acs.jpcc.1c07862
J. Phys. Chem. C 2021, 125, 24429−24439
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

changing the time for washing. By using this procedure, we EDA or EDTA, used for protecting Rh species during
obtained Na-Rh@S-1 catalysts with Rh contents in a range of hydrothermal synthesis could not affect the size of Rh
0.73−0.77 wt % and Na contents in a range of 0.13−0.26 wt % particles. It is noteworthy that the average micropore size of
(Table 1). S-1 zeolite with the MFI structure is 0.55 nm, significantly
smaller than the size of Rh nanoparticles. Therefore, the Rh
Table 1. Some Physicochemical Properties and nanoparticles in the Rh@S-1 and Na-Rh@S-1 catalysts are not
Compositions of Synthesized Catalysts encapsulated in the micropores of S-1 but are tightly
surrounded and fixed by the framework of S-1.33−36 The
content mean size of Rh particles in the reference sample (i.e., the
(wt %)a
0.17Na-Rh/S-1 prepared by the impregnation) was 4.2 nm
surface pore CO2
area volume chemisorption (Figure S4), larger than those in the Na-Rh@S-1 catalysts.
sample (m2 g−1) (cm3 g−1) Na Rh (μmol g−1)b XPS measurements were carried out to gain insights into the
S-1 360 0.27 state of surface Rh species. It is of interest that the XPS signal
Rh@S-1 326 0.23 0.77 21.9 of Rh 3d could not be discerned for the 0.19Na-Rh@S-1
0.13Na-Rh@S-1 302 0.21 0.13 0.76 23.7 catalyst, whereas the corresponding signal could be clearly
0.19Na-Rh@S-1 304 0.22 0.19 0.73 32.3 observed for the reference 0.17Na-Rh/S-1 sample (Figure S5).
0.26Na-Rh@S-1 303 0.21 0.26 0.74 24.1 This phenomenon is consistent with our expectation that the
0.17Na-Rh/S-1 295 0.22 0.17 0.75 26.6 Rh nanoparticles in the 0.19Na-Rh@S-1 catalyst are embedded
a
Determined by ICP-OES measurement. bMeasured by a magnetic in S-1 zeolite. To provide further evidence about the location
suspension balance at 298 K. of Rh species, we conducted the tomographic TEM character-
izations of the 0.19Na-Rh@S-1 and Rh@S-1 catalysts. The
The XRD measurements showed that the synthesized resin-embedded catalyst was cut into 60 nm slices by using a
samples had the typical diffraction patterns of the well- diamond knife to expose the internal view. By this technique,
crystallized MFI structure of S-1 zeolite (Figure S1), suggesting we confirmed that almost all of the Rh nanoparticles were
that our method of introducing Rh complex prior to uniformly embedded within S-1 (Figures 2c and S6a),
hydrothermal synthesis did not obstruct the crystallization of consistent with our previous result for the Pd@S-1 catalyst
S-1. No diffraction peaks belonging Rh2O3 could be observed synthesized by the same method.38 In contrast, in the 0.17Na-
in our catalysts probably because of the low Rh content or the Rh/S-1 sample, the Rh nanoparticles were mainly located on
small size of Rh species. The N2 physisorption showed that the the outer surface of S-1 (Figures S4 and S5). The dark-field
surface areas and pore volumes of our catalysts were 290−330 STEM and EDS mapping results for the 0.19Na-Rh@S-1 and
m2 g−1 and 0.21−0.27 cm3 g−1, respectively (Table 1 and Rh@S-1 catalysts provided evidence for the uniform dispersion
Figure S2). The surface area and pore volume for the catalysts of Rh throughout the S-1 crystals (Figures 2d−h and S6b−e).
were somewhat lower than those of the pristine S-1. Furthermore, the dark-field TEM, EDS mapping, and line-scan
The SEM micrographs showed that the catalysts were in elemental analysis results showed that Rh and Na were almost
hexagonal morphology and the average sizes of crystals were at identical locations over the 0.19Na-Rh@S-1 catalyst
230−260 nm (Figures 2a and S3). From the TEM micro- (Figures 2g−h and S7), suggesting that Na+ was in close
graphs, it is clear that the Rh particles were uniformly dispersed contact with Rh nanoparticles.
throughout the S-1 crystals (Figures 2b and S4). The mean 3.2. Catalytic Behaviors for Hydrogenation of CO2
sizes of Rh particles in the Rh@S-1 and Na-Rh@S-1 catalysts and CO. We investigated the catalytic behaviors of the Rh@S-
with different Na contents were close to each other and were 1 and Na-Rh@S-1 catalysts for CO2 hydrogenation. The Rh@
in the range of 2.4−2.6 nm, suggesting that the ligand, either S-1 catalyst showed a CO2 conversion of 2.9% and CH4 was

Figure 2. Electron microscopy results of the 0.19Na-Rh@S-1 catalyst. (a) SEM images and size distributions. (b) TEM image and size distributions
of Rh nanoparticles. (c) Tomographic TEM image. (d−h) Dark-field STEM image and the corresponding elemental mappings for Si, O, Rh, and
Na elements.

24432 https://doi.org/10.1021/acs.jpcc.1c07862
J. Phys. Chem. C 2021, 125, 24429−24439
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 3. Catalytic behaviors of Rh-based catalysts for CO2 hydrogenation. (a) CO2 conversion and product selectivity. (b) STY of C2H5OH and
the fraction of C2H5OH in C2+ oxygenates. Reaction conditions: catalyst loading = 0.15 g; H2/CO2 = 3; P = 5 MPa; T = 523 K; F = 15 mL min−1,
time on stream = 20 h. C2+OH: C2+ oxygenates.

formed as the major product with a small amount of CH3OH


(Figure 3a). No ethanol or other C2+ oxygenates was detected
over this catalyst without Na+ modifier. The Na-Rh@S-1
catalysts with Na + promoters showed enhanced CO 2
conversions and product distributions different from those
for the Rh@S-1 catalyst. The selectivity of CH4 decreased and
those of CH3OH and C2H5OH increased significantly. The
formation of CO via the RWGS reaction also occurred in the
presence of Na (Table S1). There is an optimum Na content,
and the 0.19Na-Rh@S-1 catalyst with a Na content of 0.19 wt
% or a Na/Rh molar ratio close to 1 demonstrated the best Figure 4. Catalytic behaviors of Rh-based catalysts for CO
performance for the formation of C2H5OH. Thus, the hydrogenation. The insets represent the distributions of C2+
modification of Rh sites by excessive Na+ is unbeneficial to oxygenates. (a) Rh@S-1. (b) 0.19Na-Rh@S-1. (c) 0.17Na-Rh/S-1.
Reaction conditions: catalyst loading = 0.15 g; H2/CO = 2; P = 5
the CO2 hydrogenation.22 The CO2 conversion and C2H5OH MPa; T = 523 K; F = 15 mL min−1, time on stream = 20 h. C2+oxy:
selectivity reached 10 and 24%, respectively, over the 0.19Na- C2+ oxygenates. C2+HC: C2+ hydrocarbons. MA: methyl acetate. EA:
Rh@S-1 catalyst, both of which were the highest among the ethyl acetate.
Na-Rh@S-1 catalysts synthesized in this work. The space-time
yield (STY) of C2H5OH reached 72 mmol gRh−1 h−1 over the
18%. It is noteworthy that the selectivity of C2H5OH was very
0.19Na-Rh@S-1 catalyst (Figure 3b), which outperformed not low, and instead, acetaldehyde, acetic acid, methyl acetate, and
only other catalysts synthesized in the present work but also ethyl acetate dominated the C2+ oxygenates (Figure S8b).
the Rh-based catalysts reported previously (Table S2). Thus, the Rh@S-1 catalyst without Na+ modifier could already
In addition to CH3OH and C2H5OH, C3−C5 alcohols catalyze the C−C coupling or the chain growth to form C2+
(denoted as C3+OH) were also formed (Figure S8a), but their oxygenates and hydrocarbons in the CO hydrogenation.
selectivity was quite low (<5%). The selectivity of C2+ Unexpectedly, different from the CO2 hydrogenation, where
hydrocarbons was also very low (≤1.5%) over that of the CO2 conversion increased by Na modification, for the CO
Na-Rh@S-1 catalysts (Table S1). The fraction of C2H5OH in hydrogenation, the CO conversion rather decreased markedly
C2+ oxygenates was around 90% over the Na-Rh@S-1 catalysts from 4.1 to 1.0% after the modification of Rh@S-1 by Na
(Figure 3b). The high fraction of C2H5OH in the C2+ (Figure 4b). At the same time, the selectivity of CH4 decreased
oxygenates would simply the product separation in the future and that of CH3OH increased significantly. The shift of
practical application. Furthermore, as compared to 0.19Na- methanation to methanol formation by Na+ modification is
Rh@S-1, the reference 0.17Na-Rh/S-1 catalyst showed lower similar to that for CO2 hydrogenation. However, the
CO2 conversion and lower C2H5OH selectivity as well as selectivities of ethanol and C2+ oxygenates were not
higher CH4 selectivity (Figure 3a). The STY of C2H5OH was significantly enhanced, and that of C2+ hydrocarbons was
about 28 mmol gRh−1 h−1 (Figure 3b), 2.5 times lower as decreased by Na+ modification. These observations indicate
compared to that of the 0.19Na-Rh@S-1 catalyst. Therefore, that the Na modification also enhances CH3OH formation by
both the Na+ modification and the confinement effect play suppressing CH4 formation but does not accelerate C−C
crucial roles in the hydrogenation of CO2 to ethanol. coupling in CO hydrogenation. Therefore, the 0.19Na-Rh@S-
Three representative catalysts (i.e., Rh@S-1, 0.19Na-Rh@S- 1 catalyst becomes a methanol-synthesis catalyst in CO
1, and 0.17Na-Rh/S-1) were chosen for the CO hydrogenation hydrogenation despite the low CO conversion activity.
study to gain insights into the similarities and differences in Another difference from CO2 hydrogenation is that the
Rh-catalyzed CO and CO2 hydrogenation reactions. The Rh@ reference catalyst (i.e., 0.17Na-Rh/S-1 prepared by impregna-
S-1 catalyst showed a CO conversion of 4.1% and CH4 tion) showed higher activity and selectivity of C2+ oxygenates
selectivity of 46% (Figure 4a). Different from CO2 hydro- in CO hydrogenation than the 0.19Na-Rh@S-1 catalyst
genation, the CO hydrogenation over the Rh@S-1 catalyst (Figure 4c). The selectivity of methanol was lower over the
provided C2+ oxygenates with a considerable selectivity (27%), 0.17Na-Rh/S-1. We further carried out CO hydrogenation
and C2+ hydrocarbons were also formed with a selectivity of over 0.19Na-Rh@S-1 and 0.17Na-Rh/S-1 catalysts at higher
24433 https://doi.org/10.1021/acs.jpcc.1c07862
J. Phys. Chem. C 2021, 125, 24429−24439
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

reaction temperatures. Although CO conversions increased did not occur during the reaction. The TEM studies revealed
with an increase in temperature from 523 to 573 K, the change that the mean size of Rh nanoparticles in the 0.19Na-Rh@S-1
in product selectivity is insignificant (Table S3). The 0.19Na- catalyst changed insignificantly after the reaction (Figure
Rh@S-1 catalyst was selective for CH3OH formation, whereas S11a). However, the mean size of Rh nanoparticles over the
the 0.17Na-Rh/S-1 catalyst showed higher selectivity of C2+ 0.17Na-Rh/S-1 catalyst increased significantly from 4.2 to 7.8
oxygenates. The selectivity of C2+ oxygenates was 58% at a CO nm after the reaction (Figure S11b). These results
conversion of 4.8% at 573 K. However, unlike in the demonstrate that the rigid framework of S-1 zeolite is able to
hydrogenation of CO2, the C2+ oxygenates in CO hydro- stabilize the embedded Rh nanoparticles and prevent their
genation over the 0.17Na-Rh/S-1 catalyst were broadly aggregation under the reaction conditions, whereas the Rh
distributed, including ethanol, acetaldehyde, acetic acid, methyl particles without the confinement effect would suffer from
acetate, and ethyl acetate, and the selectivity of C2H5OH was sintering, leading to catalyst deactivation of the impregnated
only 4.8% (Table S3). catalyst. Therefore, the confinement of the active Na-modified
3.3. Stability of Na-Modified Rh Catalyst for CO2 Rh nanoparticles by the framework of zeolite is a useful
Hydrogenation and the Confinement Effect. The catalyst strategy to enhance the catalyst stability. The present work also
deactivation is a critical issue for Rh-catalyzed CO2 hydro- offers a new example of application of the metal@zeolite type
genation because of the irreversible sintering of nano- catalysts for stable catalysis.
particles.16 We have examined the long-term stability of the 3.4. Functions of the Na+ Modifier. Our present work
0.19Na-Rh@S-1 and 0.17Na-Rh/S-1 catalysts in CO2 hydro- has demonstrated that the Na modifier plays crucial roles in
genation. As displayed in Figure 5, the 0.19Na-Rh@S-1 catalyst determining the catalytic behaviors of Rh catalysts for the
hydrogenation of CO2 and CO. We studied in more depth the
roles of Na modifier in affecting the chemical state of Rh
species to gain insights into the structure−performance
relationship. The H2-TPR result showed that the reduction
of Rh species in the Rh@S-1 catalyst occurred at temperatures
below 423 K (Figure 6a). Upon the introduction of Na+ to
Rh@S-1, the reduction peak shifted to higher temperatures
and became broader or was even split into two peaks, which
might be attributed to the stepwise reduction of Rh3+ to Rh+
and then Rh+ to Rh0 or the reduction of Rh species of different
states.42,43 Therefore, the Na+ modifier suppressed the
reduction of Rh species.
Figure 5. Stability of the 0.19Na-Rh@S-1 and 0.17Na-Rh/S-1 In situ CO-adsorbed FT-IR was employed to study the
catalysts for CO2 hydrogenation. Reaction conditions: catalyst loading chemical state of Rh species over our catalysts. Prior to each
= 0.15 g; H2/CO2 = 3; P = 5 MPa; T = 523 K; F = 15 mL min−1. measurement, the catalyst was pretreated by a procedure
similar to that adopted for catalytic reactions (i.e., with a H2 (5
was stable in 100 h of reaction; both the CO2 conversion vol %)/Ar flow at 673 K for 30 min). Then, the sample was
(∼10%) and STY of ethanol (∼72 mmol gRh−1 h−1) remained cooled to room temperature in H2, followed by CO adsorption
unchanged with time on stream. However, the 0.17Na-Rh/S-1 at room temperature and then evacuation to remove gaseous
catalyst underwent deactivation in 50 h of reaction under the CO. The IR bands at ∼2045 and ∼1860 cm−1 were observed
same conditions. for the Rh@S-1 catalyst (Figure 6b), which could be assigned
We carried out characterizations of the 0.19Na-Rh@S-1 and to the linear CO and bridged CO adsorption on the Rh0 sites,
0.17Na-Rh/S-1 catalysts after reactions. The XRD and N2 respectively.44,45 This observation indicates that Rh0 dominates
physisorption results clarified that the crystallinity and porous the Rh@S-1 catalyst after the reduction treatment. Additional
structure of S-1 in the catalyst did not undergo any changes bands at ∼2080 and ∼2010 cm−1 appeared for the Na-Rh@S-1
after 100 h of reaction (Figures S9 and 10 and Table S4), catalyst (Figure 6b), and these two bands were attributable to
confirming the good stability of S-1 zeolite under our reaction the symmetric and asymmetric stretching of CO adsorbed on
conditions. The contents of Rh and Na in the 0.19Na-Rh@S-1 Rh+ sites in the form of germinal dicarbonyl (i.e., Rh-
and 0.17Na-Rh/S-1 catalysts did not change significantly after (CO)2).46,47 Therefore, the CO-adsorbed FT-IR results
reactions (Table S5), revealing that the leakage of Rh or Na provide further evidence that surface of the Rh@S-1 catalyst

Figure 6. Characterizations of Rh@S-1 and Na-Rh@S-1 catalysts. (a) H2-TPR profiles. (b) CO-adsorbed FT-IR spectra. (c) Rh 3d XPS spectra
after Ne ion sputtering.

24434 https://doi.org/10.1021/acs.jpcc.1c07862
J. Phys. Chem. C 2021, 125, 24429−24439
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

is dominated by Rh0 species, and the presence of Na+ modifier


induces the generation of Rh+ together with Rh0 species. It is
noteworthy that the intensity of adsorbed CO bands decreased
for the 0.26Na-Rh@S-1 catalyst with a Na/Rh molar ratio of
∼1.6, suggesting that some surface Rh sites were covered by
Na+ at a higher Na content.
Because Rh species were embedded in the zeolite matrix, the
conventional XPS measurement was not useful (Figure S5).
Thus, we carried out the sputtering of the surface of Rh-
embedded catalysts by Ne ions with repeated sputtering−
annealing cycles to gain information about the chemical state
of the Rh species. The sample was pretreated under H2
atmosphere at 673 K prior to sputtering and XPS measure-
ments. After sputtering treatment for 25 times, the Rh signals
could be detected and the binding energy of Rh 3d5/2 for the
Rh@S-1 catalyst was observed at 307.1 eV (Figure 6c), which
corresponded to metallic Rh species.42,43 In contrast, the Rh
3d spectrum of the 0.19Na-Rh@S-1 catalyst contained both
metallic Rh0 (307.1 eV) and oxidic Rh+ (308.2 eV)
components. Therefore, the XPS results obtained for the
samples after Ne ions sputtering offer further evidence that the
presence of Na+ retards the reduction of Rh species and both Figure 7. In situ DRIFTS spectra of adsorbed species on Rh@S-1 and
Rh0 and Rh+ coexist on the surface of Na-promoted Rh 0.19Na-Rh@S-1 catalysts. (a) Rh@S-1 under 2.5 MPa (CO2 + H2) at
nanoparticles embedded with S-1 zeolite. 523 K. (b) 0.19Na-Rh@S-1 under 2.5 MPa (CO2 + H2) at 523 K. (c)
The experimental results described above all point out that Rh@S-1 after switching CO2 to H2 (2.5 MPa) at 523 K. (d) 0.19Na-
Rh0 is the predominant species over the Rh@S-1 catalyst and Rh@S-1 after switching CO2 to H2 (2.5 MPa) at 523 K.
that a fraction of oxidic Rh+ species coexist on the Na-modified
Rh@S-1 catalyst due to role of Na+. We further measured the 2720, 1742, 1605, and 1362 cm−1 attributable to HCOO*
amount of CO2 chemisorption, and the result revealed that the species and those at 2815 and 1042 cm−1 assignable to CH3O*
presence of Na+ promoted the chemisorption of CO2 (Table species were observed.52−54 The HCOO*, CH3O*, and CO*
1). The enhanced CO2 adsorption and the presence of Rh+ species might be formed by the hydrogenation of adsorbed
species after Na+ modification may also decrease the H2 CO3* species.49,55 Moreover, the IR bands at 2962, 2874, and
adsorption and dissociation ability,20,22 thus inhibiting the 1092 cm−1, which could be attributed to CH3CH2O*
formation of hydrocarbons, in particular CH4, and promoting species,52 were observed. This species is proposed to be the
the C−O bond activation and CO insertion for C−C coupling key adsorbed species for C2H5OH formation.
over Rh sites. Therefore, the increased amount of CO2 To further gain information about the evolution of surface
chemisorption and the coexistence of Rh0 and Rh+ species in adsorbed species in the reaction course of CO2 hydrogenation,
the presence of Na+ may mainly account for the higher CO2 we carried out in situ DFIFTS studies by first introducing CO2
conversion and the larger ethanol selectivity for CO 2 (2.5 MPa) into the IR cell and then switching the feed from
hydrogenation. CO2 to H2 (2.5 MPa) at 523 K. The result for the Rh@S-1
3.5. Possible Reaction Mechanism. We carried out in catalyst showed that the decrease in the intensities of the bands
situ DRIFTS studies over the Rh@S-1 catalyst at 2.5 MPa and belonging to adsorbed bidentate CO3* species and the increase
523 K to gain insights into the adsorbed species and possible in those of the bands belonging to CH4 (3015 cm−1) and
reaction intermediates. Under (H2 + CO2) atmosphere, IR CHx* species (2956, 2926, and 2856 cm−1) (Figure 7c). The
bands at 1300, 1260, and 1200 cm−1 were observed over the intensity of adsorbed CO* species kept strong at the initial
Rh@S-1 catalyst, and these bands could be assigned to the stage, implying that the reduction of bidentate CO3* species to
adsorbed bidentate carbonate (CO3*) species (Figure 7a).48,49 CO is more facile than the hydrogenation of CO intermediate.
At the same time, the IR bands at 2052 and 1850 cm−1, which For the 0.19Na-Rh@S-1 catalyst, the intensity of IR bands of
were attributable to the linear- and bridge-adsorbed CO bidentate CO3* species similarly decreased significantly after
species,44 and the bands at 2923 and 2852 cm−1 assignable to switching to H2. However, the intensities for the IR bands of
stretching vibrations of CHx species50,51 were observed over HCOO* species (2720, 1605, and 1362 cm−1) and CH3O*
the Rh@S-1 catalyst. The IR band at 3016 cm−1 belonging to species (2815 cm−1) increased with time. This suggests that
gaseous CH4 also appeared. The intensities of IR bands the HCOO* and CH3O* species are generated from the
attributable to CH4 and adsorbed CHx species increased with hydrogenation of bidentate CO3* species.49 The bands at
time on stream. These observations are consistent with our 2962, 2874, and 1090 cm−1 assignable to the CH3CH2O*
catalytic result that the CO2 hydrogenation over Rh@S-1 species emerged and intensified with time. The adsorbed CO
mainly offers CH4 product and further suggest that the species (1910 and 1820 cm−1) appeared instantly after the
formation of CH4 proceeds through CO intermediate. switching to H2 (Figure 7d), and their intensities decreased to
The same in situ DRIFTS studies for the 0.19Na-Rh@S-1 some extent with time. These observations are consistent with
catalyst under 2.5 MPa (H2 + CO2) and 523 K also showed the mechanism that the CH3CH2O* species is generated by
the presence of the adsorbed bidentate CO3* species (IR the insertion of CO.
bands at 1294 and 1204 cm−1) and CO species (IR bands at The reaction mechanisms for CO2 hydrogenation on the
1930 and 1874 cm−1, Figure 7b). Moreover, the IR bands at Rh@S-1 and Na-Rh@S-1 catalysts are discussed based on the
24435 https://doi.org/10.1021/acs.jpcc.1c07862
J. Phys. Chem. C 2021, 125, 24429−24439
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

in situ DFIFTS results and the information about the active contributing to higher CO2 conversion activity. Second, the
site. We first briefly discuss the mechanism for CO 2 zeolite confinement effect can restrict the sintering of Na-
hydrogenation over the Rh@S-1 catalyst without Na + modified Rh nanoparticles, making the Na-Rh@S-1 catalyst
modification. Previous studies on CO hydrogenation suggest highly stable as compared to the reference Na-Rh/S-1 catalyst.
that the Rh0 particles mainly functions for the dissociation of Furthermore, both Rh0 and Rh+ species coexist over the Na-
CO as well as H2, favoring the formation of CH4 as the major Rh@S-1 catalyst with Na+ modification, which play a crucial
product.56,57 As already mentioned, the Rh@S-1 catalyst role in the coupling between adsorbed CHx and CO species to
contains mainly Rh0 sites. The in situ DRIFTS studies for form ethanol, and the confinement by zeolite may stabilize the
CO2 hydrogenation showed that the adsorbed CO and CHx Rh0−Rh+ pair site during the reaction. Therefore, the synergy
species were mainly present on the Rh@S-1 surface. In of the confinement effect of S-1 zeolite and the Rh0−Rh+ pair
agreement with our experimental observation that CH4 is the site boosts the formation of ethanol by CO2 hydrogenation.
predominant product during the CO2 hydrogenation over the It is well-known that Rh-based catalysts are efficient for the
Rh@S-1 catalyst, we propose that CO is the intermediate for conversion of syngas to C2+ oxygenates.30,31 It is quite
the formation of CH4 from CO2 over this catalyst. That is, unexpected that our 0.19Na-Rh@S-1 catalyst, which is active
CO2 is first hydrogenated to CO adsorbed on Rh0 surfaces,
and selective for the hydrogenation of CO2 to ethanol,
which undergoes further dissociation and hydrogenation to
exhibited much lower CO conversion and ethanol selectivity
adsorbed CHx species and CH4.
during CO hydrogenation even at an increased temperature
Next, we discuss the CO2-hydrogenation mechanism on the
Na+-modified Rh@S-1 catalyst. It is accepted that the Rh+ site (Figure 4 and Table S3). We speculate that unlike in CO2
cannot dissociate CO, favoring the formation of CH3OH as the hydrogenation, where the formed CO intermediate is in low
major product during CO hydrogenation.56 The Rh0−Rh+ pair concentration, the high concentration of CO in CO hydro-
site is believed to be necessary for the formation of ethanol, genation may poison the Rh0 site,58,59 in particular over our
which is believed to proceed by the insertion of adsorbed CO catalyst with small sizes of confined Rh particles, which only
species into Rh−CH3 or the coupling between nondissociative contain a fraction of Rh0. In fact, the Na-Rh@S-1 catalyst
CO and dissociative species.39,42 Our characterizations have shows high selectivity of CH3OH during CO hydrogenation
pointed out that Rh0 and Rh+ coexist on the Na-Rh@S-1 despite the low CO conversion. This also implies that the
catalyst. The in situ DRIFTS studies revealed that in addition major active site over this catalyst is Rh+, which cannot
to the adsorbed CO and bidentate CO3* species, the adsorbed dissociate CO molecules.
HCOO*, CH3O*, and CH3CH2O* species also appeared on
the catalyst surface. Therefore, we propose a possible 4. CONCLUSIONS
mechanism for CO2 hydrogenation on the Na-Rh@S-1 catalyst
in Figure 8. In brief, CO2 is first adsorbed on the catalyst We have successfully synthesized Na-modified Rh nano-
particles embedded in zeolite S-1 (Na-Rh@S-1) by a ligand-
protected crystallization method. The Rh content is around
0.75 wt %, and the Na content can be regulated in a range of
0.13−0.26 wt %. The mean size of the Rh particles in close
contact with Na+ is around 2.5 nm. The 0.19Na-Rh@S-1
catalyst is efficient for CO2 hydrogenation to ethanol, offering
a space−time yield of ethanol as high as 72 mmol gRh−1 h−1,
which outperforms those obtained over Rh-based catalysts
reported to date. The presence of Na+ induces the coexistence
of Rh0 and Rh+ species, dramatically shifting the major product
from methane to oxygenates mainly including methanol and
ethanol. The smaller size of Rh particles and the enhancement
in CO2 adsorption by Na+ contribute to improving the CO2
conversion. The confinement by the rigid framework of S-1
protects the Rh nanoparticles from aggregation, offering
superior stability of the Na-Rh@S-1 catalyst during long-
term CO2 hydrogenation. We further found that compared to
CO2 hydrogenation the 0.19Na-Rh@S-1 catalyst shows
Figure 8. Proposed reaction mechanism for CO2 hydrogenation over significantly lower activity during the CO hydrogenation and
Na-Rh@S-1 catalysts. that the catalyst is more selective toward the formation of
methanol instead of ethanol.


surface in the form of *CO3 species, which is hydrogenated by
the adsorbed H (Ha) species to CO* or HCOO*. The ASSOCIATED CONTENT
HCOO* species is further hydrogenated to CH3O* and
CH3OH on the Rh+ site. The CO* species undergo * Supporting Information

dissociation and hydrogenation to the CHx* species on the The Supporting Information is available free of charge at
Rh0 site, which may either be converted to CH4 or undergo https://pubs.acs.org/doi/10.1021/acs.jpcc.1c07862.
coupling with CO* to form the CHxCO*, which is further
hydrogenated into CH3CH2OH via CH3CH2O*. Zeolite is XRD, N2 adsorption−desorption isotherms, SEM micro-
believed to play the following roles. First, the confinement by graphs, TEM micrographs, XPS spectra, ICP-OES, GC
S-1 zeolite offers Rh nanoparticles with smaller sizes, graphs and additional catalytic data (PDF)

24436 https://doi.org/10.1021/acs.jpcc.1c07862
J. Phys. Chem. C 2021, 125, 24429−24439
The Journal of Physical Chemistry C


pubs.acs.org/JPCC Article

AUTHOR INFORMATION https://pubs.acs.org/10.1021/acs.jpcc.1c07862


Corresponding Authors
Jincan Kang − State Key Laboratory of Physical Chemistry of Author Contributions

Solid Surfaces, Innovation Laboratory for Sciences and F.Z. and W.Z. contributed equally to this work.
Technologies of Energy Materials of Fujian Province, Notes
National Engineering Laboratory for Green Chemical The authors declare no competing financial interest.


Productions of Alcohols, Ethers and Esters, College of
Chemistry and Chemical Engineering, Xiamen University,
Xiamen 361005, PR China; Email: kangjc@xmu.edu.cn ACKNOWLEDGMENTS
Ye Wang − State Key Laboratory of Physical Chemistry of This work was supported by the National Key Research and
Solid Surfaces, Innovation Laboratory for Sciences and Development Program of Ministry of Science and Technology
Technologies of Energy Materials of Fujian Province, of China (No. 2019YFE0104400), and the National Natural
National Engineering Laboratory for Green Chemical Science Foundation of China (Nos. 22121001, 91945301,
Productions of Alcohols, Ethers and Esters, College of 21872112, 21972116, and 22172123).


Chemistry and Chemical Engineering, Xiamen University,
Xiamen 361005, PR China; orcid.org/0000-0003-0764- REFERENCES
2279; Email: wangye@xmu.edu.cn
(1) Sanz-Perez, E. S.; Murdock, C. R.; Didas, S. A.; Jones, C. W.
Authors Direct Capture of CO2 from Ambient Air. Chem. Rev. 2016, 116,
Fuyong Zhang − State Key Laboratory of Physical Chemistry 11840−11876.
of Solid Surfaces, Innovation Laboratory for Sciences and (2) Olah, G. A.; Prakash, G. K.; Goeppert, A. Anthropogenic
Chemical Carbon Cycle for a Sustainable Future. J. Am. Chem. Soc.
Technologies of Energy Materials of Fujian Province, 2011, 133, 12881−12898.
National Engineering Laboratory for Green Chemical (3) Shaner, M. R.; Atwater, H. A.; Lewis, N. S.; McFarland, E. W. A
Productions of Alcohols, Ethers and Esters, College of Comparative Technoeconomic Analysis of Renewable Hydrogen
Chemistry and Chemical Engineering, Xiamen University, Production Using Solar Energy. Energy Environ. Sci. 2016, 9, 2354−
Xiamen 361005, PR China 2371.
Wei Zhou − State Key Laboratory of Physical Chemistry of (4) Centi, G.; Quadrelli, E. A.; Perathoner, S. Catalysis for CO2
Solid Surfaces, Innovation Laboratory for Sciences and Conversion: A Key Technology for Rapid Introduction of Renewable
Technologies of Energy Materials of Fujian Province, Energy in the Value Chain of Chemical Industries. Energy Environ. Sci.
National Engineering Laboratory for Green Chemical 2013, 6, 1711−1731.
Productions of Alcohols, Ethers and Esters, College of (5) Shih, C. F.; Zhang, T.; Li, J.; Bai, C. Powering the Future with
Liquid Sunshine. Joule 2018, 2, 1925−1949.
Chemistry and Chemical Engineering, Xiamen University, (6) Zhou, W.; Cheng, K.; Kang, J.; Zhou, C.; Subramanian, V.;
Xiamen 361005, PR China Zhang, Q.; Wang, Y. New Horizon in C1 Chemistry: Breaking the
Xuewei Xiong − State Key Laboratory of Physical Chemistry Selectivity Limitation in Transformation of Syngas and Hydro-
of Solid Surfaces, Innovation Laboratory for Sciences and genation of CO2 into Hydrocarbon Chemicals and Fuels. Chem. Soc.
Technologies of Energy Materials of Fujian Province, Rev. 2019, 48, 3193−3228.
National Engineering Laboratory for Green Chemical (7) Jiang, X.; Nie, X.; Guo, X.; Song, X.; Chen, J. G. Recent
Productions of Alcohols, Ethers and Esters, College of Advances in Carbon Dioxide Hydorgenation to Methanol via
Chemistry and Chemical Engineering, Xiamen University, Heterogeneous Catalysis. Chem. Rev. 2020, 120, 7984−8034.
Xiamen 361005, PR China (8) Zhong, J.; Yang, X.; Wu, Z.; Liang, B.; Huang, Y.; Zhang, T.
Yuhao Wang − State Key Laboratory of Physical Chemistry of State of the Art and Perspectives in Heterogeneous Catalysis of CO2
Hydrogenation to Methanol. Chem. Soc. Rev. 2020, 49, 1385−1413.
Solid Surfaces, Innovation Laboratory for Sciences and (9) Asare Bediako, B. B.; Qian, Q.; Han, B. Synthesis of C2+
Technologies of Energy Materials of Fujian Province, Chemicals from CO2 and H2 via C-C Bond Formation. Acc. Chem.
National Engineering Laboratory for Green Chemical Res. 2021, 54, 2467−2476.
Productions of Alcohols, Ethers and Esters, College of (10) Ma, Z.; Porosoff, M. D. Development of Tandem Catalysts for
Chemistry and Chemical Engineering, Xiamen University, CO2 Hydrogenation to Olefins. ACS Catal. 2019, 9, 2639−2656.
Xiamen 361005, PR China (11) Wang, D.; Xie, Z.; Porosoff, M. D.; Chen, J. G. Recent
Kang Cheng − State Key Laboratory of Physical Chemistry of Advances in Carbon Dioxide Hydrogenation to Produce Olefins and
Solid Surfaces, Innovation Laboratory for Sciences and Aromatics. Chem. 2021, 7, 2277−2311.
Technologies of Energy Materials of Fujian Province, (12) Nezam, I.; Zhou, W.; Gusmão, G. S.; Realff, M. J.; Wang, Y.;
National Engineering Laboratory for Green Chemical Medford, A. J.; Jones, C. W. Direct Aromatization of CO2 via
Combined CO2 Hydrogenation and Zeolite-based Acid Catalysis. J.
Productions of Alcohols, Ethers and Esters, College of CO2 Util. 2021, 45, 101405.
Chemistry and Chemical Engineering, Xiamen University, (13) Gao, P.; Zhang, L.; Li, S.; Zhou, Z.; Sun, Y. Novel
Xiamen 361005, PR China; orcid.org/0000-0002-7112- Heterogeneous Catalysts for CO2 Hydrogenation to Liquid Fuels.
4700 ACS Cent. Sci. 2020, 6, 1657−1670.
Qinghong Zhang − State Key Laboratory of Physical (14) Pang, J. F.; Zheng, M. Y.; Zhang, T. Synthesis of Ethanol and
Chemistry of Solid Surfaces, Innovation Laboratory for Its Catalytic Conversion. Adv. Catal. 2019, 64, 89−191.
Sciences and Technologies of Energy Materials of Fujian (15) Surisetty, V. R.; Dalai, A. K.; Kozinski, J. Alcohols as Alternative
Province, National Engineering Laboratory for Green Fuels: An Overview. Appl. Catal., A 2011, 404, 1−11.
Chemical Productions of Alcohols, Ethers and Esters, College (16) Xu, D.; Wang, Y.; Ding, M.; Hong, X.; Liu, G.; Tsang, S. C. E.
of Chemistry and Chemical Engineering, Xiamen University, Advances in Higher Alcohol Synthesis from CO2 Hydrogenation.
Chem. 2021, 7, 849−881.
Xiamen 361005, PR China (17) Zeng, F.; Mebrahtu, C.; Xi, X.; Liao, L.; Ren, J.; Xie, J.; Heeres,
Complete contact information is available at: H. J.; Palkovits, R. Catalysis Design for Higher Alcohols Synthesis by

24437 https://doi.org/10.1021/acs.jpcc.1c07862
J. Phys. Chem. C 2021, 125, 24429−24439
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

CO2 Hydrogenation: Trends and Future Perspectives. Appl. Catal., B (39) Wang, C.; Zhang, J.; Qin, G.; Wang, L.; Zuidema, E.; Yang, Q.;
2021, 291, 120073. Dang, S.; Yang, C.; Xiao, J.; Meng, X.; et al. Direct Conversion of
(18) Zhang, S.; Wu, Z.; Liu, X.; Hua, K.; Shao, Z.; Wei, B.; Huang, Syngas to Ethanol within Zeolite Crystals. Chem. 2020, 6, 646−657.
C.; Wang, H.; Sun, Y. A Short Review of Recent Advances in Direct (40) Zecevic, J.; Vanbutsele, G.; de Jong, K. P.; Martens, J. A.
CO2 Hydrogneation to Alcohols. Top. Catal. 2021, 64, 371−394. Nanoscale Intimacy in Bifunctional Catalysts for Selective Conversion
(19) Nie, X.; Li, W.; Jiang, X.; Guo, X.; Song, C. Recent Advances in of Hydrocarbons. Nature 2015, 528, 245−248.
Catalytic CO2 Hydrogenation to Alcohols and Hydrocarbons. Adv. (41) Cheng, K.; van der Wal, L. I.; Yoshida, H.; Oenema, J.; Harmel,
Catal. 2019, 65, 121−233. J.; Zhang, Z.; Sunley, G.; Zečević, J.; de Jong, K. P. Impact of the
(20) Kusama, H.; Okabe, K.; Sayama, K.; Arakawa, H. CO2 Spatial Organization of Bifunctional Metal-Zeolite Catalysts for
Hydrogenation to Ethanol over Promoted Rh/SiO2 Catalysts. Catal. Hydroisomerization of Light Alkanes. Angew. Chem., Int. Ed. 2020,
Today 1996, 28, 261−266. 59, 3592−3600.
(21) Kusama, H.; Okabe, K.; Sayama, K.; Arakawa, H. Ethanol (42) Sheerin, E.; Reddy, G. K.; Smirniotis, P. Evaluation of Rh/
Synthesis by Catalytic Hydrogenation of CO2 over Rh-Fe/SiO2 CexTi1‑xO2 Catalysts for Synthesis of Oxygenates from Syngas Using
Catalysts. Energy 1997, 22, 343−348. XPS and TPR Techniques. Catal. Today 2016, 263, 75−83.
(22) Kitamura Bando, K.; Soga, K.; Kunimori, K.; Arakawa, H. Effect (43) Chen, Y.; Zhang, H.; Ma, H.; Qian, W.; Jin, F.; Ying, W. Direct
of Li Additive on CO2 Hydrogenation Reactivity of Zeolite Supported Conversion of Syngas to Ethanol over Rh−Fe/γ-Al2O3 Catalyst:
Rh Catalysts. Appl. Catal., A 1998, 175, 67−81. Promotion Effect of Li. Catal. Lett. 2018, 148, 691−698.
(23) Yang, C.; Mu, R.; Wang, G.; Song, J.; Tian, H.; Zhao, Z. J.; (44) Matsubu, J. C.; Yang, V. N.; Christopher, P. Isolated Metal
Gong, J. Hydroxyl-Mediated Ethanol Selectivity of CO2 Hydro- Active Site Concentration and Stability Control Catalytic CO2
genation. Chem. Sci. 2019, 10, 3161−3167. Reduction Selectivity. J. Am. Chem. Soc. 2015, 137, 3076−3084.
(24) Wang, G.; Luo, R.; Yang, C.; Song, J.; Xiong, C.; Tian, H.; (45) Yang, N.; Yoo, J. S.; Schumann, J.; Bothra, P.; Singh, J. A.;
Zhao, Z.-J.; Mu, R.; Gong, J. Active Sites in CO2 Hydrogenation over Valle, E.; Abild-Pedersen, F.; Nørskov, J. K.; Bent, S. F. Rh-MnO
Confined VOx-Rh Catalysts. Sci. China: Chem. 2019, 62, 1710−1719. Interface Sites Formed by Atomic Layer Deposition Promote Syngas
(25) Ding, L.; Shi, T.; Gu, J.; Cui, Y.; Zhang, Z.; Yang, C.; Chen, T.; Conversion to Higher Oxygenates. ACS Catal. 2017, 7, 5746−5757.
Lin, M.; Wang, P.; Xue, N.; et al. CO2 Hydrogenation to Ethanol over (46) Schwartz, V.; Campos, A.; Egbebi, A.; Spivey, J. J.; Overbury, S.
Cu@Na-Beta. Chem. 2020, 6, 2673−2689. H. EXAFS and FT-IR Characterization of Mn and Li Promoted
(26) An, B.; Li, Z.; Song, Y.; Zhang, J.; Zeng, L.; Wang, C.; Lin, W. Titania-Supported Rh Catalysts for CO Hydrogenation. ACS Catal.
Cooperative Copper Centers in a Metal-Organic Framework for 2011, 1, 1298−1306.
Selective Conversion of CO2 to Ethanol. Nat. Catal. 2019, 2, 709− (47) Heyl, D.; Rodemerck, U.; Bentrup, U. Mechanistic Study of
717. Low-Temperature CO2 Hydrogenation over Modified Rh/Al2O3
(27) Wang, L.; Wang, L.; Zhang, J.; Liu, X.; Wang, H.; Zhang, W.; Catalysts. ACS Catal. 2016, 6, 6275−6284.
Yang, Q.; Ma, J.; Dong, X.; Yoo, S. J.; et al. Selective Hydrogenation (48) Du, H.; Williams, C. T.; Ebner, A. D.; Ritter, J. A. In Situ FTIR
of CO2 to Ethanol over Cobalt Catalysts. Angew. Chem., Int. Ed. 2018, Spectroscopic Analysis of Carbonate Transformations during
57, 6104−6108. Adsorption and Desorption of CO2 in K-Promoted HTlc. Chem.
(28) Zhang, S.; Liu, X.; Shao, Z.; Wang, H.; Sun, Y. Direct CO2 Mater. 2010, 22, 3519−3526.
Hydrogenation to Ethanol over Supported Co2C Catalysts: Studies on (49) Xu, D.; Ding, M.; Hong, X.; Liu, G.; Tsang, S. C. E. Selective
Support Effects and Mechanism. J. Catal. 2020, 382, 86−96. C2+ Alcohol Synthesis from Direct CO2 Hydrogenation over a Cs-
(29) Chen, Y.; Choi, S.; Thompson, L. T. Low Temperature CO2 Promoted Cu-Fe-Zn Catalyst. ACS Catal. 2020, 10, 5250−5260.
Hydrogenation to Alcohols and Hydrocarbons over Mo2C Supported (50) Wang, C.; Guan, E.; Wang, L.; Chu, X.; Wu, Z.; Zhang, J.;
Metal Catalysts. J. Catal. 2016, 343, 147−156. Yang, Z.; Jiang, Y.; Zhang, L.; Meng, X.; et al. Product Selectivity
(30) Luk, H. T.; Mondelli, C.; Ferré, D. C.; Stewart, J. A.; Pérez- Controlled by Nanoporous Environments in Zeolite Crystals
Ramírez, J. Status and Prospects in Higher Alcohols Synthesis from Enveloping Rhodium Nanoparticle Catalysts for CO2 Hydrogenation.
Syngas. Chem. Soc. Rev. 2017, 46, 1358−1426. J. Am. Chem. Soc. 2019, 141, 8482−8488.
(31) Spivey, J. J.; Egbebi, A. Heterogeneous Catalytic Synthesis of (51) Wang, Y.; Kattel, S.; Gao, W.; Li, K.; Liu, P.; Chen, J. G.; Wang,
Ethanol from Biomass-derived Syngas. Chem. Soc. Rev. 2007, 36, H. Exploring the Ternary Interactions in Cu-ZnO-ZrO2 Catalysts for
1514−1528. Efficient CO2 Hydrogenation to Methanol. Nat. Commun. 2019, 10,
(32) Gogate, M. R.; Davis, R. J. Comparative Study of CO and CO2 1166.
Hydrogenation over Supported Rh−Fe Catalysts. Catal. Commun. (52) Xu, D.; Ding, M.; Hong, X.; Liu, G. Mechanistic Aspects of the
2010, 11, 901−906. Role of K Promotion on Cu−Fe-Based Catalysts for Higher Alcohol
(33) Wang, N.; Sun, Q.; Yu, J. Ultrasmall Metal Nanoparticles Synthesis from CO2 Hydrogenation. ACS Catal. 2020, 10, 14516−
Confined within Crystalline Nanoporous Materials: A Fascinating 14526.
Class of Nanocatalysts. Adv. Mater. 2019, 31, 1803966. (53) Borchert, H.; Jürgens, B.; Zielasek, V.; Rupprechter, G.;
(34) Wang, H.; Wang, L.; Xiao, F. Metal@Zeolite Hybrid Materials Giorgio, S.; Henry, C. R.; Bäumer, M. Pd Nanoparticles with Highly
for Catalysis. ACS Cent. Sci. 2020, 6, 1685−1697. Defined Structure on MgO as Model Catalysts: An FTIR Study of the
(35) Li, Y.; Yu, J. Emerging Applications of Zeolites in Catalysis, Interaction with CO, O2, and H2 Under Ambient Conditions. J. Catal.
Separation and Host-Guest Assembly. Nat. Rev. Mater. 2021, 2007, 247, 145−154.
DOI: 10.1038/s41578-021-00347-3. (54) Yang, R.; Fu, Y.; Zhang, Y.; Tsubaki, N. In Situ DRIFT Study of
(36) Wang, Y.; Wang, C.; Wang, L.; Wang, L.; Xiao, F. Zeolite Fixed Low-Temperature Methanol Synthesis Mechanism on Cu/ZnO
Metal Nanoparticles: New Perspective in Catalysis. Acc. Chem. Res. Catalysts from CO2-Containing Syngas Using Ethanol Promoter. J.
2021, 54, 2579−2590. Catal. 2004, 228, 23−35.
(37) Sun, Q.; Wang, N.; Fan, Q.; Zeng, L.; Mayoral, A.; Miao, S.; (55) Kattel, S.; Liu, P.; Chen, J. G. Tuning Selectivity of CO2
Yang, R.; Jiang, Z.; Zhou, W.; Zhang, J.; et al. Subnanometer Hydrogenation Reactions at the Metal/Oxide Interface. J. Am. Chem.
Bimetallic Platinum-Zinc Clusters in Zeolites for Propane Dehydro- Soc. 2017, 139, 9739−9754.
genation. Angew. Chem. 2020, 132, 19618−19627. (56) Kawai, M.; Uda, M.; Ichikawa, M. The Electronic State of
(38) Wang, W.; Zhou, W.; Li, W.; Xiong, X.; Wang, Y.; Cheng, K.; Supported Rh Catalysts and the Selectivity for the Hydrogenation of
Kang, J.; Zhang, Q.; Wang, Y. In-Situ Confinement of Ultrasmall Carbon Monoxide. J. Phys. Chem. 1985, 89, 1654−1656.
Palladium Nanoparticles in Silicalite-1 for Methane Combustion with (57) Mo, X.; Gao, J.; Umnajkaseam, N.; Goodwin, J. G., Jr La, V,
Excellent Activity and Hydrothermal Stability. Appl. Catal., B 2020, and Fe Promotion of Rh/SiO2 for CO Hydrogenation: Effect on
276, 119142. Adsorption and Reaction. J. Catal. 2009, 267, 167−176.

24438 https://doi.org/10.1021/acs.jpcc.1c07862
J. Phys. Chem. C 2021, 125, 24429−24439
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

(58) Choi, Y.; Liu, P. Mechanism of Ethanol Synthesis from Syngas


on Rh(111). J. Am. Chem. Soc. 2009, 131, 13054−13061.
(59) Wang, Y.; Luo, H.; Liang, D.; Bao, X. Different Mechanisms for
the Formation of Acetaldehyde and Ethanol on the Rh−Based
Catalysts. J. Catal. 2000, 196, 46−55.

24439 https://doi.org/10.1021/acs.jpcc.1c07862
J. Phys. Chem. C 2021, 125, 24429−24439

You might also like