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

pubs.acs.

org/JACS Article

Wettability Engineering of Solar Methanol Synthesis


Zhe Lu,# Yangfan Xu,# Zeshu Zhang,# Junchuan Sun, Xue Ding, Wei Sun, Wenguang Tu, Yong Zhou,
Yingfang Yao,* Geoffrey A. Ozin,* Lu Wang,* and Zhigang Zou
Cite This: J. Am. Chem. Soc. 2023, 145, 26052−26060 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: Engineering the wettability of surfaces with hydrophobic


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

organics has myriad applications in heterogeneous catalysis and the large-scale


chemical industry; however, the mechanisms behind may surpass the proverbial
Downloaded via MAE FAH LUANG UNIV on April 15, 2024 at 21:34:14 (UTC).

hydrophobic kinetic benefits. Herein, the well-studied In2O3 methanol synthesis


photocatalyst has been used as an archetype platform for a hydrophobic
treatment to enhance its performance. With this strategy, the modified samples
facilitated the tuning of a wide range of methanol production rates and
selectivity, which were optimized at 1436 μmol gcat−1 h−1 and 61%, respectively.
Based on in situ DRIFTS and temperature-programmed desorption-mass
spectrometry, the surface-decorated alkylsilane coating on In2O3 not only
kinetically enhanced the methanol synthesis by repelling the produced polar
molecules but also donated surface active H to facilitate the subsequent
hydrogenation reaction. Such a wettability design strategy seems to have
universal applicability, judged by its success with other CO2 hydrogenation catalysts, including Fe2O3, CeO2, ZrO2, and Co3O4.
Based on the discovered kinetic and mechanistic benefits, the enhanced hydrogenation ability enabled by hydrophobic alkyl groups
unleashes the potential of the surface organic chemistry modification strategy for other important catalytic hydrogenation reactions.

■ INTRODUCTION
The conversion of CO2 with green H2 into valuable methanol
sites.9−12 We hypothesize that an effective means to circum-
vent both the selectivity and durability problem is to make the
has been envisioned as a promising approach to alleviate the catalyst surface hydrophobic by decorating it with chemically
atmospheric CO2 concentration and curtail the reliance on tethered alkyl chains.9,10 By this means, it is envisioned that an
fossil fuels.1,2 Conventional methanol synthesis is energy- organically tailored hydrophobic catalyst surface will selectively
intensive, necessitating high reaction temperatures (>240 °C) repel strongly polar water and methanol relative to weakly
and pressures (∼5 MPa).3,4 Incorporating solar energy via polar CO products, thereby favoring the methanol reaction
photothermal catalysis has the potential to surmount this pathway over CO.10,13 Despite the advantage in the desorption
obstacle;5−8 however, exploring the efficient photothermal of produced molecules, other mechanistic benefits remain to
catalysts with stable performance remains challenging. be discovered. Thus, it would be of great scientific interest and
In detail, the synthesis of methanol (eq 1) in the technological importance if the hydrophobic surface could play
heterogeneous catalytic hydrogenation of CO2 with respect other roles to facilitate the catalytic process.
to the competing reverse water gas shift (RWGS) CO pathway Herein, the viability of the surface wettability effect was
(eq 2) is pressure- and temperature-dependent. The methanol experimentally demonstrated for a series of metal oxide
and CO pathways are respectively exothermic and endother- catalysts, exemplified by indium, iron, zirconium, cerium, and
mic, and by the thermodynamic principle of Le Chatelier for cobalt oxides. Each of these is observed to positively shift the
these equilibrium reactions, the former is favored at low
kinetic rate and thermodynamic conversion toward methanol
temperatures and high pressures, while the latter is at high
temperatures with little sensitivity to pressure.1−3 or CH4 compared to CO. This advance portends a universal
paradigm for activity and selectivity management in gas-phase
CO2 + 3H 2 CH3OH + H 2O H0 = 50 kJ mol 1
heterogeneous CO2 photocatalysis.
(1)
Received: July 10, 2023
CO2 + H 2 CO + H 2O H 0= + 39 kJ mol 1
Revised: November 6, 2023
(2) Accepted: November 7, 2023
Both reactions, however, are challenged by the cogeneration Published: November 20, 2023
of product water as it can deactivate the catalyst by
coordinating to and subsequently restructuring surface active

© 2023 American Chemical Society https://doi.org/10.1021/jacs.3c07349


26052 J. Am. Chem. Soc. 2023, 145, 26052−26060
Journal of the American Chemical Society pubs.acs.org/JACS Article

■ EXPERIMENTAL SECTION
Synthesis of Samples. The precipitation method was used to
was heated (10 °C min−1) under argon. The effluent gas was analyzed
using a PFEIFFER OmniSTAR. Similarly, CO2-TPD and water-TPD
were tested through the same processes with the preadsorption gas of
synthesize In2O3. To be specific, 2 g of In(NO3)3·xH2O (99.99%) was CO2 and 3% H2O/Ar, respectively. The only difference for water-
dissolved in 200 mL of deionized water. Then, under stirring, 100 mL TPD is that it was tested directly after preadsorption without a
of a 0.1 M Na2CO3·10H2O (99.99%) solution was added until the pH sweeping procedure. Moreover, we conducted in situ diffuse
reached 8. The mixture was aged at 80 °C under 3 h continuous reflectance infrared Fourier transform (DRIFTS) to characterize the
stirring and then collected and washed by suction filtration. After adsorbed species on the sample surface and further monitor the
drying at 80 °C overnight, the sample was calcined at 400 °C (2 °C catalytic reaction process in real-time. The tests were performed on a
min−1) for 1 h to obtain In2O3 (denoted as S0). Co3O4 was prepared Thermo Nicolet iS50 FTIR spectrometer with a high-precision
by the same method. The preparation of alkylsilane-modified In2O3 mercury−cadmium−telluride (MCT) detector cooled by liquid
was achieved by silanization reactions (Figure S1). 100 mg portion of nitrogen. Beyond an infrared spectrometer, the experimental system
prepared In2O3 was mixed with 2 mL of toluene. Subsequently, 10 μL consists of a diffuse reflectance accessory (Praying Mantis, Harrick), a
of triethoxyoctylsilane (TEOOS, 98%) was added. The mixture was high-temperature cell (HVC, Harrick, equipped with CaF2 windows),
then refluxed at 50 °C for 3 h under stirring. After that, the products a gas system, and a heating and temperature control device. Before the
were centrifugally separated and washed with toluene and ethanol in situ tests, the catalyst was purged under a flowing Ar atmosphere at
several times. Then, they were transferred into a vacuum oven at 120 373 K for 3 h to remove possible impurity molecules. Then, the
°C and dried for 24 h. The obtained catalysts were named S2. background spectrum under an Ar atmosphere at different temper-
Similarly, the catalysts synthesized from 5, 40, and 120 μL of TEOOS atures was collected to subtract the interference of H2O and CO2 on
were named S1, S3, and S4, respectively. Alkylsilane-decorated ZrO2 the sample spectrum. The scanning resolution was 2 cm−1, and the
(99.99%), CeO2 (99.9%), Fe2O3 (99.8%), and Co3O4 were number of scans was 32 times. In the H2 adsorption experiment, the
synthesized by the same process as S2. The products were named ratio of H2 and Ar was 6:14. The ratio of CO2 and Ar was 2:18 in CO2
ZrO2−Si, CeO2−Si, Fe2O3−Si, and Co3O4−Si, respectively. C1, C4, adsorption tests. After testing at 250 °C under CO2/Ar, H2 was
C6, C10, and C12 with the same theoretical Si content as S2 were introduced into the system (H2/CO2/Ar: 6/2/12). The residual gas
prepared using a similar process, with the feedstock of methyl- after D2 adsorption was detected via a Hiden HPR-20 EGA mass
triethoxysilane (98%), buthyltriethoxysilane (94%), hexyltrimethox- spectrometer.
ysilane (98%), decyltriethoxysilane (98%), and dodecyltriethoxysilane Evaluation of Catalytic Performance. The catalytic perform-
(97%), respectively. ance was obtained through a batch reactor with a silica viewport at the
Structural Characterizations. The Si content was determined top (Figure S2). The reactor (25.2 mL) was loaded with 5 mg of
through an inductively coupled plasma spectrometer (ICP) using an catalysts and purged with H2 (4.5 bar) and CO2 (1.5 bar) at room
Agilent 5110. The water droplet contact angles were obtained with a temperature. The reaction was conducted at 200 °C for 0.5 h. A Xe
Sindin SDC-500 contact angle meter. The attenuated total reflectance light (320−780 nm) was used for providing light irradiation from top
infrared spectroscopy (ATR-IR) tests were performed on a Thermo to bottom with a light intensity of ∼12 suns. Products were analyzed
Nicolet iS50 Fourier-transform infrared (FTIR) spectrometer with an with an online gas chromatograph (GC, Panna A91Plus) equipped
integrated ATR module. The phase compositions of the samples were with a flame-ionized detector (FID) and a thermal conductivity
characterized by a Rigaku Smartlab diffractometer equipped with a Cu detector (TCD). The product after isotope labeling reactions were
Kα radiation source (λ = 1.54056 Å). The X-ray source was operated analyzed through a Xevo G2-XS QTof mass spectrometer using the
at 40 kV and 200 mA. The phase identification was made by atmospheric pressure ionization (API) mode. The selectivity and
comparison to the Joint Committee on Powder Diffraction Standards reaction rate of product i (i = methanol or CO, etc.) were calculated
(JCPDS). The N2 physical adsorption−desorption test can be used to as follows
analyze the specific surface area and pore structure. The test was ni
confirmed on a Micromeritics ASAP 2010 instrument. The specific ri =
surface area was analyzed by the Brunauer−Emmett−Teller (BET) (mcat × t )
method. The pore structure was investigated by the Barrett−Joyner−
Halenda (BJH) method. The surface compositional and chemical Si = ri / ri×100%
states of catalysts were analyzed through X-ray photoelectron i
spectroscopy (XPS) using an AXIS SUPRA+ spectrometer with an
Al Kα X-ray source. The operating conditions were chosen as 15 kV where ni is the moles of product i, t is the reaction time, and mcat is the
and 27 A. The carbon C 1s at 284.8 eV was used for calibration. catalyst sample weight.
Electron spin resonance (ESR) spectra of the samples were also
recorded for more complete and fine structural information. It was
conducted at room temperature using a Bruker A300 EPR
■ RESULTS AND DISCUSSION
Structural Characterization. The hydrophobic In2O3
spectrometer operated at the X-band frequency. The morphology of
samples were prepared via a silanization reaction under reflux
samples was investigated through transmission electron microscopy
(TEM). The TEM images, high-resolution TEM (HRTEM) images, with toluene. The original In2O3 is labeled as S0. From the ICP
and EDX mapping were obtained using a JEOL JEM-2100F results, the samples with Si contents of 0.050, 0.085, 0.103, and
microscope at the accelerating voltage of 200 kV. The weight loss 0.117 wt % are labeled as S1, S2, S3, and S4, respectively.
and heat changes of samples at different temperatures were During the thermal treatment under N2, all samples exhibited
investigated through thermogravimetry analysis (TGA) with differ- weight loss and prominent derivative thermogravimetry
ential scanning calorimetry (DSC) and mass spectrometry (MS). The (DTG) peaks with similar DSC curves from 300 to 400 °C
samples were tested in a Netzsch STA449F3-QMS403D thermogravi- (Figure S3). It implies that the decorated alkyl chains are
metric analyzer under N2. thermally stable until 300 °C in N2. The excellent thermal
Mechanism Study. H2 temperature-programmed desorption stability can be inferred from the minor weight loss of S1−S4
(TPD) was conducted on a chemisorption analyzer (Micromeritics
Autochem II 2920). 100 mg portion of the as-prepared sample was
from 50 to 200 °C. As shown in Figures 1A and S4A, the
pretreated under flowing argon for 0.5 h at room temperature and increased alkylsilane content leads to an increase in contact
then heated to 300 °C for 0.5 h and further 400 °C for 1 h. Then, the angle: 23.7, 57.5, 137.7, 147.6, and 156.0° from S0 to S4,
sample was cooled to 50 °C under flowing argon. After stabilization, a respectively. In addition, the S1 to S4 samples show a receding
mixture gas of 10% H2 in N2 was introduced for 1 h. Subsequently, water signal from 3700 to 3500 cm−1 (Figure S4B).14 It implies
the gas was switched into argon for 1 h sweeping. Then, the sample the reinforced hydrophobicity on S1−S4. From powder X-ray
26053 https://doi.org/10.1021/jacs.3c07349
J. Am. Chem. Soc. 2023, 145, 26052−26060
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 1. Structural and elemental characterizations of S1−S4. The contact angle (A), XRD spectra (B), and high-resolution XPS spectra (C) of In
3d over S0 and S2. (D) The scanning TEM and energy-dispersive X-ray spectroscopy mapping results of S2. (E) Schematic of the proposed surface
structure of S1−S4 (region I: highly dense surface with crystallized/ordered structure; region II: less-dense surface with disordered structure).

diffraction (XRD) results (Figure 1B), all samples exhibit a crystallinity of the decorated alkylsilanes is increased ranging
cubic bixbyite structured phase (#PDF 06-0416),8 and SiO2- from S1 to S4 (Figure S4C).
related signals cannot be observed, indicating the decorated Photothermal Catalytic Performance. The catalytic
alkylsilane coating is amorphous or in extremely minimal performance for photothermal CO2 hydrogenation is shown
content. in Figure 2A. The optimum alkylsilane coating over S2 leads to
The Si signals can be found in the XPS of S1−S4, indicating excellent photothermal catalytic activity. At 200 °C, S0, the
the successful modification (Figure S5). In the high-resolution ordinary In2O3, exhibits the methanol rate of 189 and 709
In 3d core-level spectra (Figure 1C), two peaks at 444 and 452 μmol gcat−1 h−1 for both dark and light conditions with
eV can be assigned to In 3d5/2 and 3d3/2, respectively.15,16 It methanol selectivity of 50 and 53%, respectively. On the
can be found that the In 3d peaks gradually move to higher contrary, S2 exhibits a methanol rate of 242 and 1436 μmol
binding energy with an ∼0.1 eV shift from S0 to S4, which gcat−1 h−1 with methanol selectivities of 60 and 61% under dark
could be attributed to the generation of electron-deficient In and light, respectively. The comparison of the methanol rate,
by the added alkylsilanes.16,17 Owing to the higher electro- selectivity, and turnover frequency (TOF) of S2 under light
negativity of the added Si and O (Si−O−In bonds),18 the with other photothermal catalysts under similar conditions is
shown in Tables S1 and S2. The relatively low estimated TOF
outer electrons of In could be attracted and result in lowered
of our samples (the S2 TOF of 4.7) could be caused by the
electron density. Furthermore, an electric field can be
high density of active sites compared to other materials based
generated due to the different electronegativity and thus
on metal active sites.1,5,13 Under light conditions, the methanol
enhance polarization and result in enhanced ESR signals with rate of S2 is twice as high as that of S0. In sharp contrast, with
smaller g values at 2.003 (from 2.004 of S0) over S1−S4 the increasing amount of surface alkyl chains, S3 and S4 exhibit
(Figure S6).19,20 lower activity, with the methanol rate of 635 and 570 μmol
All samples exhibit a similar appearance, with a particle size gcat−1 h−1 under light, implying that the formed Si−O−In
of ∼10 nm (Figure S7). According to the elemental mapping connection may block active sites and decrease the catalytic
shown in Figure 1D, the Si element is uniformly distributed on performance.
the In2O3 nanoparticles (Figure 1D).14,21 The decoration of The stability of S0 and S2 was further studied and is shown
alkylsilanes generates mesopores with a radius of ∼2 nm, in Figure 2B. The methanol rate and selectivity of S2 can be
leading to enlarged specific surface areas from 72 m2 g−1 of S0 maintained at 1313 μmol gcat−1 h−1 and 55% for 20 runs,
to ∼90 m2 g−1 of S1−S4 (Figures S8C and S9).10 Based on the whereas the methanol rate and selectivity of S0 keep
aforementioned results, a representative schematic of the decreasing gradually to 334 μmol gcat−1 h−1 and 36% in the
decorated In2O3 surface is proposed in Figure 1E. Different 20th test. The post-S2 sample exhibits a similar surface Si
amounts of decorated alkylsilanes induce different crystal- amount (0.0402 wt %) to the fresh S2 (0.0396 wt %) via XPS,
linities and coverage of alkyl chains.22,23 The order of with the In2O3 (222) facet and a lattice distance of 0.29 nm
26054 https://doi.org/10.1021/jacs.3c07349
J. Am. Chem. Soc. 2023, 145, 26052−26060
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 2. Catalytic performance. (A) Production rates and methanol selectivity of S0−S4. The error bars are based on the standard deviation. (B)
Stability tests under the light of S0 and S2. (C) Mass spectrometry of the 13CO2 labeling reaction of S2. (D) The corresponding Arrhenius plot
under light conditions on S0 and S2. (E) Dependence of formation rates on the H2 partial pressure of S0 and S2. Test conditions: ∼12 suns, 6 bar,
H2/CO2: 3:1 (purge with 0.5−3 bar Ar in E), 200 °C (A−C, E).

and remains hydrophobic without aggregation (Figures S8 and a thermal insulating layer, effectively curbing thermal
S10 and Table S3). According to the in situ DRIFTS conduction and augmenting the local temperature of S2 by
measurements and water-contained activity tests, S2 exhibits approximately 12 °C.28,29 Nonetheless, even when compared
a stable Si−O−Si signal (1157 cm−1)24 under 0.24% H2O, to the S0 sample tested in the dark at 242 °C, the S2 sample
which is much higher than the water content of ∼0.09% under tested at 200 °C under light still exhibits a significantly higher
reaction circumstance, as shown in Figures S11 and S12. methanol production rate, with an increase of ∼464 μmol
Water-TPD with MS was further conducted to confirm the gcat−1 h−1 (Figure S14C). Judging by these results, the
stability with water steam (Figure S13). Methane (m/z = 16) decorated alkylsilanes may yield distinctive mechanistic
cannot be detected from S2 before 420 °C, indicating that the advantages in the CO2 hydrogenation.
alkyl chain retains its chemical stability until 420 °C with Furthermore, the light also plays a vital role in selectively
water. The minor physically desorbed water from 50 to 200 °C lowering the methanol activation energy by 19.0 and 24.7 kJ
on S2 suggests a robust ability to promote water draining. mol−1 on S0 and S2, respectively. Compared to the
Thus, the aggregation-induced deactivation from water can be performance tested without light under the same local
ameliorated, improving reaction stability.25 The 13CO2-labeling temperature, the methanol rate is enhanced (from 675.3 to
results indicate that all products from S2 are generated by the 709.0 μmol gcat−1 h−1), and the selectivity increases (from 50.8
reaction gas (Figure 2C). to 52.7%) when light is irradiated onto S0, suggesting the
The activation energy of S0 and S2 was estimated under involvement of a photochemical pathway in methanol
illumination and is shown in Figure 2D. The S2 sample production (Figure S14D).15 S2 exhibits a more substantial
exhibits a much higher activation energy for CO (90.9 kJ photochemical effect, with a methanol rate enhancement of
mol−1) than that of methanol (55.7 kJ mol−1), whereas the 403.6 μmol gcat−1 h−1 and an increased methanol selectivity
activation energy of methanol and CO of S0 are 67.3 kJ mol−1 from 50.3 to 60.5% (Figure S14D). As the light intensities
and 87.1 kJ mol−1, respectively. Without illumination, the increase by 5 suns, a remarkable enhancement in methanol rate
methanol activation energy of S2 is 80.4 kJ mol−1, which is is observed on S2, from 797 to 1763 μmol gcat−1 h−1 (Figure
lower than that of S0 by ∼6 kJ mol−1 (Figure S14A,B). The S14E). This enhancement surpasses that of CO (from 352 to
assessment of local temperature was conducted by employing 1039 μmol of gcat−1 h−1). One can deduce that the light may
parallel Arrhenius plots of CO under both dark and light facilitate intermediate stabilization30,31 and active hydrogen
conditions, the light-induced local temperature of S0 and S2 at transfer,32,33 further promoting CO2 hydrogenation to
200 °C was estimated to be approximately 229 and 241 °C, methanol. To further confirm the mechanistic benefits of
respectively (Figure S14A).8 The generation of heat arises alkylsilanes, the dependence of the H2 partial pressure on the
from the nonradiative relaxation of light-induced charge methanol rate was investigated. As shown in Figure 2E, the S2
carriers on In2O3.26,27 The presence of alkylsilanes establishes sample shows a much lower order of H2 at 0.23 for methanol
26055 https://doi.org/10.1021/jacs.3c07349
J. Am. Chem. Soc. 2023, 145, 26052−26060
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 3. Functions of alkyl chains. In situ DRIFTS spectra in the hydride region (A) and hydroxyl region (B) at 300 °C under Ar over S2. (C)
TGA and MS result over S0 and S2. The TPD-MS results of the absorption gas of CO2 (D) and H2 (E) over S2 and S0. (F) In situ DRIFTS spectra
in the hydroxyl region at 75 °C under H2 (Ar/H2: 14/6) over S2 [α: CH3; β: CH2; δ: In−H−; γ: InH2−; κ: In−OH+; the arrow direction in (A−F)
means the longer test time]. (G) The catalytic performance of Fe2O3, CeO2, ZrO2, and Co3O4 and their samples after alkylsilane modification
(labeled with “Si-”). Test conditions: ∼12 suns, 6 bar, H2/CO2: 3:1, 200 °C.

formation, whereas the H2 order of S0 for methanol formation surface of S2 (Figure S16). Thus, one can deduce that the alkyl
is ∼0.61. This can be further confirmed by the result of chains over the catalyst are the potential H source by
decreased CO2 order for CO formation over S2 (Figure S15) dehydrogenation with the nearby indium site.34,37 In sharp
and indicates an enriched surface H concentration on S2. contrast, S0 exhibits none of the above phenomena (Figures
Mechanistic Study. Kinetically speaking, the hydrophobic S16 and S17). To further confirm this idea, TGA-MS was used
coating can efficiently remove the polar molecules, such as the to verify whether the H2 could be released with thermal
produced water, methanol, and CO molecules, right shift the treatment. As can be seen from Figure 3C, only the decorated
reaction equilibrium, and result in better overall catalytic S2 (thermally stable up to 350 °C) could generate H2 (m/z =
performance. However, such repulsion should theoretically 2) with a peak at 300 °C, indicating that the H atoms from
benefit methanol formation, a product with a more vital alkyl chains could also participate in the reaction as the active
polarity than CO. Thus, the methanol selectivity should be H. The corresponding CO2-TPD-MS of S2 indicates that the
improved with increased hydrophobicity. However, as shown CO could be produced by consuming the H atom from alkyl
in Figure 2A, ranging from S1 to S4, a volcano-shaped chains, and no CO could be observed from S0 (Figure 3D). It
methanol selectivity plot could be observed, implying that an agrees well with the previous deduction.
external mechanistic role could be provided by the hydro- Only regenerable species could be considered as a valuable
phobic coating. contribution to the catalytic process. The corresponding H2-
To study the mechanistic role of the alkyl chains, detailed in TPD-MS of S2 indicates a superior ability to activate H2 than
situ DRIFTS experiments were performed over S0 and S2. S0 (Figure 3E). In the meantime, H2 could also be generated
Interestingly, the indium hydride-related species can be during the TPD process for S2, and only water could be
observed at 1426 and 1640 cm−1 (In−H−), 1574 cm−1 produced from S0. Another in situ DRIFTS measurement was
(InH2−), as well as 3554 and 3650 cm−1 (In−OH+) under conducted in the H2 atmosphere to verify whether methyl and
an Ar atmosphere (Figure 3A and B).34−36 On the other hand, methylene could be regenerated. As can be seen from Figure
the negative signals of methyl at 2968 cm−1 and methylene at 3F, the rise of peaks from 2970 to 2860 cm−1 represents the
2924 and 2854 cm−1, indicate a decrement in the amount regeneration of methyl and methylene groups, and the
(Figure 3B).23 The introduction of CO2 can form bicarbonates presence of a hydride signal at 1574 cm−1 agrees well with
(1635 cm−1),4 also showing the presence of active H on the the one formed under the Ar atmosphere (Figures 3F and
26056 https://doi.org/10.1021/jacs.3c07349
J. Am. Chem. Soc. 2023, 145, 26052−26060
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 4. Mechanism study. (A) The catalytic performance is regulated by regulating the length of carbon chains. (B) In situ DRIFTS spectra in the
C−O stretch region at 250 °C under the reaction gas (CO2/H2/Ar: 2/6/12) over S2. (C) The schematic of the proposed reaction mechanism over
S2 (σ: CO; τ: HOCO*; ω: CH3O*; the arrow direction means the longer test time).

S18). To further confirm this regeneration property, D2 was and CH3O* (1115 cm−1)40 can be observed. According to
used as the control experiment, and the formation of CH2D previous studies, HOCO* could be the intermediate source for
and CHD2 at 2933 cm−1, the electron ionization (EI) mass further hydrogenation and CO formation.8,41,42 The proposed
spectra of H, and the formation of CHxD3‑xOD with D2 and reaction mechanism can be concluded as shown in Figure 4C,
CO2 can be observed (Figures S19−S21). The H-D exchange and the CO2 hydrogenation to methanol could be conducted
proves the participation of carbon chains in H2 activation. This through the CO-RWGS process.8,40−42 The absorbed CO2 can
function of alkyl chains is also effective in Fe2O3, CeO2, ZrO2, be converted into HOCO* (iv), which can be directly
and Co3O4. As shown in Figure 3G, the production of hydrogenated to HCO* (v) with sufficient H species. The
methanol, CH4, and C2H6 is enhanced after alkylsilane HOCO* species can also be hydrogenated and decomposed to
decoration. It further confirms that the alkyl chains could CO and water (dashed box in Figure 4C). The fewer H species
donate external H to the reaction environment and shift the over S0 may result in a higher possibility of CO formation.
reaction from RWGS to deep hydrogenated products, such as Due to the more vital ability to repel water (Figures S4, S12,
methanol, CH4, and C2H6 (Figures S22−S25).38 and S13), generated water from HOCO* hydrogenation on S2
As shown in Figure 4A, the number of H atoms depends on can be removed simultaneously. The introduction of light
the length of the alkyl chains, and various chain lengths result further promotes these processes on S2 by capitalizing on the
in different catalytic performances. While there is no significant synergistic interplay of three factors. (1) The energy barrier for
change in the CO rate (∼960 μmol gcat−1 h−1) from C1 to C8, methanol production in excited states is reduced, employing a
the methanol rate increases monotonically from 792 to 1436 mechanism akin to that of S0.15 (2) Light promotes the
μmol gcat−1 h−1 with an increased selectivity from 45 to 61%. transfer of active hydrogen species.32,33 (3) Light-induced
The stable CO rate and improved methanol rate could be localized electrons on oxygen vacancies can be transferred to
caused by H donation from the alkyl chains. The poor catalytic the antibonding orbitals of CO2 molecules, activating and
performance from short-chain samples is mainly due to the stabilizing intermediates,30,31,43 ultimately facilitating further
coverage of alkylsilane and the absence of H donation (Figure hydrogenation on S2. Therefore, the S2 sample can effectively
S26). While the chain length is increased from C8 to C12, the convert the CO2 into methanol.
hydrophobicity keeps increasing (Figure S27) and favors the
kinetically fast RWGS reaction rather than methanol8,38 and
ends up with improved overall catalytic performance and
■ CONCLUSIONS
In conclusion, the decorated alkylsilane coatings could not
enhanced CO selectivity. only kinetically enhance the reaction by repelling the produced
To further study the reaction mechanism of S2, the in situ polar molecules but also donate surface active H to facilitate
DRIFTS spectra under the reaction gas (CO2/H2 = 1:3) at 250 the subsequent hydrogenation reaction. Such a wettability
°C were recorded and shown in Figure 4B. The signals of CO engineering strategy can also be applied to other hydro-
(2222 cm−1),15 HOCO* (1520, 1385, and 1253 cm−1),39,40 genation catalysts, including Fe2O3, CeO2, ZrO2, and Co3O4.
26057 https://doi.org/10.1021/jacs.3c07349
J. Am. Chem. Soc. 2023, 145, 26052−26060
Journal of the American Chemical Society pubs.acs.org/JACS Article

The enhanced hydrogenation ability from the corresponding Collaborative Innovation Center of Advanced
alkyl chains unleashes the potential of surface organic Microstructures, School of Physics, Nanjing University,
decoration strategy for other important hydrogenation catalysis Nanjing 210093, P. R. China; orcid.org/0000-0002-
reactions. 9480-2586
Zhigang Zou − School of Science and Engineering, The
■ ASSOCIATED CONTENT
* Supporting Information

Chinese University of Hong Kong, Shenzhen, Shenzhen
518172, P. R. China; National Laboratory of Solid State
The Supporting Information is available free of charge at Microstructures, Collaborative Innovation Center of
https://pubs.acs.org/doi/10.1021/jacs.3c07349. Advanced Microstructures, School of Physics, Nanjing
University, Nanjing 210093, P. R. China; orcid.org/
Catalyst performance data; TGA-DSC, contact angle, 0000-0003-2092-8335
ATR-IR, XPS, ESR, TEM, N2 adsorption−desorption,
XRD, water-TPD-MS, MS, and in situ DRIFTS Complete contact information is available at:
characterization data; schematic of the proposed https://pubs.acs.org/10.1021/jacs.3c07349
salinization process; diagram of the reactor; comparison
Author Contributions
of reported photothermal catalytic performance; com- #
Z.L., Y.X., and Z.Z. contributed equally to this paper.
parison of reported TOF for methanol synthesis; surface
Si ratio from XPS; CO2 chemisorption quantity; Notes
calculation of water content and TOF; and references The authors declare no competing financial interest.
(PDF)
■ ACKNOWLEDGMENTS

■ AUTHOR INFORMATION
Corresponding Authors
All authors appreciate the support of the National Key R&D
Program of China (grants no. 2021YFF0502000), the National
Natural Science Foundation of China (grants no. 52102311
Yingfang Yao − School of Science and Engineering, The and 22208270), UofT-ZJU Joint seed Fund, the Program for
Chinese University of Hong Kong, Shenzhen, Shenzhen Guangdong Introducing Innovative and Entrepreneurial
518172, P. R. China; National Laboratory of Solid State Teams (grants no. 2019ZT08L101), the Special Fund for
Microstructures, Collaborative Innovation Center of the Sci-tech Innovation Strategy of Guangdong Province
Advanced Microstructures, School of Physics, Nanjing (grants no. 210629095860472), the Shenzhen Natural Science
University, Nanjing 210093, P. R. China; orcid.org/ Foundation (grants no. GXWD20201231105722002-
0000-0003-4823-0094; Email: yaoyingfang@nju.edu.cn 20200824163747001), the Shenzhen Key Laboratory of Eco-
Geoffrey A. Ozin − Solar Fuels Group, Department of materials and Renewable Energy (grants no.
Chemistry, University of Toronto, Toronto, Ontario M5S ZDSYS20200922160400001), NSF of Jiangsu Province (grants
3H6, Canada; orcid.org/0000-0002-6315-0925; no. BK20220006), support of Shenzhen Natural Science
Email: g.ozin@utoronto.ca Foundation 2022, and support of Provincial Talent Plan and
Lu Wang − School of Science and Engineering, The Chinese the University Development Fund (grants no. UDF01001721).
University of Hong Kong, Shenzhen, Shenzhen 518172, P. R.
China; orcid.org/0000-0002-4165-4022;
Email: lwang@cuhk.edu.cn
■ REFERENCES
(1) Gao, W.; Liang, S.; Wang, R.; Jiang, Q.; Zhang, Y.; Zheng, Q.;
Xie, B.; Toe, C. Y.; Zhu, X.; Wang, J.; et al. Industrial carbon dioxide
Authors capture and utilization: state of the art and future challenges. Chem.
Zhe Lu − School of Science and Engineering, The Chinese Soc. Rev. 2020, 49, 8584−8686.
University of Hong Kong, Shenzhen, Shenzhen 518172, P. R. (2) Tountas, A. A.; Peng, X.; Tavasoli, A. V.; Duchesne, P. N.;
China; orcid.org/0000-0002-9254-7518 Dingle, T. L.; Dong, Y.; Hurtado, L.; Mohan, A.; Sun, W.; Ulmer, U.;
Yangfan Xu − Solar Fuels Group, Department of Chemistry, et al. Towards solar methanol: past, present, and future. Adv. Sci.
University of Toronto, Toronto, Ontario M5S 3H6, Canada; 2019, 6, 1801903.
orcid.org/0000-0002-4479-6157 (3) Porosoff, M. D.; Yan, B.; Chen, J. G. Catalytic reduction of CO2
Zeshu Zhang − Ganjiang Innovation Academy, Chinese by H2 for synthesis of CO, methanol and hydrocarbons: challenges
and opportunities. Energy Environ. Sci. 2016, 9, 62−73.
Academy of Sciences, Ganzhou 341000, P. R. China (4) Xie, B.; Wong, R. J.; Tan, T. H.; Higham, M.; Gibson, E. K.;
Junchuan Sun − School of Science and Engineering, The Decarolis, D.; Callison, J.; Aguey-Zinsou, K.-F.; Bowker, M.; Catlow,
Chinese University of Hong Kong, Shenzhen, Shenzhen C. R. A.; et al. Synergistic ultraviolet and visible light photo-activation
518172, P. R. China; orcid.org/0009-0009-7663-1259 enables intensified low-temperature methanol synthesis over copper/
Xue Ding − School of Science and Engineering, The Chinese zinc oxide/alumina. Nat. Commun. 2020, 11, 1615.
University of Hong Kong, Shenzhen, Shenzhen 518172, P. R. (5) Zhao, J.; Liu, J.; Li, Z.; Wang, K.; Shi, R.; Wang, P.; Wang, Q.;
China Waterhouse, G. I. N.; Wen, X.; Zhang, T. Ruthenium-cobalt single
Wei Sun − State Key Laboratory of Silicon Materials, School of atom alloy for CO photo-hydrogenation to liquid fuels at ambient
Materials Science and Engineering, Zhejiang University, pressures. Nat. Commun. 2023, 14, 1909.
Hangzhou 310027, P. R. China (6) Zhao, J.; Zhang, T. Storage and release of active sites for solar
methanol at ambient pressure. Chem Catal. 2023, 3, 100535.
Wenguang Tu − School of Science and Engineering, The (7) Xu, Y.-F.; Tountas, A. A.; Song, R.; Ye, J.; Wei, J.-H.; Ji, S.; Zhao,
Chinese University of Hong Kong, Shenzhen, Shenzhen L.; Zhou, W.; Chen, J.-H.; Zhao, G.; et al. Equilibrium photo-
518172, P. R. China; orcid.org/0000-0001-8206-6166 thermodynamics enables a sustainable methanol synthesis. Joule 2023,
Yong Zhou − School of Science and Engineering, The Chinese 7, 738−752.
University of Hong Kong, Shenzhen, Shenzhen 518172, P. R. (8) Zhang, Z.; Mao, C.; Meira, D. M.; Duchesne, P. N.; Tountas, A.
China; National Laboratory of Solid State Microstructures, A.; Li, Z.; Qiu, C.; Tang, S.; Song, R.; Ding, X.; et al. New black

26058 https://doi.org/10.1021/jacs.3c07349
J. Am. Chem. Soc. 2023, 145, 26052−26060
Journal of the American Chemical Society pubs.acs.org/JACS Article

indium oxide�tandem photothermal CO2-H2 methanol selective (26) Mateo, D.; Albero, J.; García, H. Titanium-perovskite-
catalyst. Nat. Commun. 2022, 13, 1512−1518. supported RuO2 nanoparticles for photocatalytic CO2 methanation.
(9) Tan, M.; Tian, S.; Zhang, T.; Wang, K.; Xiao, L.; Liang, J.; Ma, Joule 2019, 3, 1949−1962.
Q.; Yang, G.; Tsubaki, N.; Tan, Y. Probing hydrophobization of a Cu/ (27) Sun, J.; Sun, W.; Wang, L.; Ozin, G. A. Uniting heat and light in
ZnO catalyst for suppression of water-gas shift reaction in syngas heterogeneous CO2 photocatalysis: optochemical materials and
conversion. ACS Catal. 2021, 11, 4633−4643. reactor engineering. Acc. Mater. Res. 2022, 3, 1260−1271.
(10) Xu, Y.; Li, X.; Gao, J.; Wang, J.; Ma, G.; Wen, X.; Yang, Y.; Li, (28) Cai, M.; Wu, Z.; Li, Z.; Wang, L.; Sun, W.; Tountas, A. A.; Li,
Y.; Ding, M. A hydrophobic FeMn@Si catalyst increases olefins from C.; Wang, S.; Feng, K.; Xu, A.-B.; et al. Greenhouse-inspired supra-
syngas by suppressing C1 by-products. Science 2021, 371, 610−613. photothermal CO2 catalysis. Nat. Energy 2021, 6, 807−814.
(11) Fang, W.; Wang, C.; Liu, Z.; Wang, L.; Liu, L.; Li, H.; Xu, S.; (29) Han, X.; Besteiro, L. V.; Koh, C. S. L.; Lee, H. K.; Phang, I. Y.;
Zheng, A.; Qin, X.; Liu, L.; et al. Physical mixing of a catalyst and a Phan-Quang, G. C.; Ng, J. Y.; Sim, H. Y. F.; Lay, C. L.; Govorov, A.;
hydrophobic polymer promotes CO hydrogenation through dehy- et al. Intensifying heat using MOF-isolated graphene for solar-driven
dration. Science 2022, 377, 406−410. seawater desalination at 98% solar-to-thermal efficiency. Adv. Funct.
(12) Zhao, J.; Li, Z.; Wang, P.; Miao, P.; Shi, R.; Waterhouse, G. I.; Mater. 2021, 31, 2008904.
Zhang, T. Hydrophobic modification for CO photo-hydrogenation to (30) Qi, Y.; Song, L.; Ouyang, S.; Liang, X.; Ning, S.; Zhang, Q.; Ye,
olefins with low CO2 selectivity. Nano Energy 2023, 110, 108350. J. Photoinduced defect engineering: enhanced photothermal catalytic
(13) Wang, H.; Liu, G.; Chen, C.; Tu, W.; Lu, Y.; Wu, S.; O’Hare, performance of 2D black In2O3‑x nanosheets with bifunctional oxygen
D.; Xu, R. Single-Ni sites embedded in multilayer nitrogen-doped vacancies. Adv. Mater. 2020, 32, 1903915.
graphene derived from amino-functionalized MOF for highly selective (31) Shao, T.; Wang, X.; Dong, H.; Liu, S.; Duan, D.; Li, Y.; Song,
CO2 electroreduction. ACS Sustain. Chem. Eng. 2021, 9, 3792−3801. P.; Jiang, H.; Hou, Z.; Gao, C.; et al. A stacked plasmonic
(14) Cargnello, M.; Jaén, J. J. D.; Garrido, J. C. H.; Bakhmutsky, K.; metamaterial with strong localized electric field enables highly
Montini, T.; Gámez, J. J. C.; Gorte, R. J.; Fornasiero, P. Exceptional efficient broadband light-driven CO2 hydrogenation. Adv. Mater.
activity for methane combustion over modular Pd@CeO2 subunits on 2022, 34, 2202367.
functionalized Al2O3. Science 2012, 337, 713−717. (32) Fang, X.; Zhang, N.; Chen, S.-C.; Luo, T. Scalable total
(15) Yan, T.; Wang, L.; Liang, Y.; Makaremi, M.; Wood, T. E.; Dai, synthesis of (−)-triptonide: serendipitous discovery of a visible-light-
Y.; Huang, B.; Ali, F. M.; Dong, Y.; Ozin, G. A. Polymorph selection promoted olefin coupling initiated by metal-catalyzed hydrogen atom
towards photocatalytic gaseous CO2 hydrogenation. Nat. Commun. transfer (MHAT). J. Am. Chem. Soc. 2022, 144, 2292−2300.
2019, 10, 2521. (33) Zhang, Z.; Tian, H.; Sun, J.; Meira, D. M.; Zhang, M.; Ding, X.;
(16) Yang, C.; Pei, C.; Luo, R.; Liu, S.; Wang, Y.; Wang, Z.; Zhao, Ji, D.; Qiu, C.; Lu, Z.; Sun, L.; et al. Light-enabled coupling of tandem
Z.; Gong, J. Strong electronic oxide support interaction over In2O3/ ethane dehydrogenation and CO2 hydrogenation. Chem Catal. 2023,
ZrO2 for highly selective CO2 hydrogenation to methanol. J. Am. 3, 100644.
(34) Wang, C.; Han, Y.; Tian, M.; Li, L.; Lin, J.; Wang, X.; Zhang, T.
Chem. Soc. 2020, 142, 19523−19531.
Main-group catalysts with atomically dispersed In sites for highly
(17) Yang, Y.; Chai, Z.; Qin, X.; Zhang, Z.; Muhetaer, A.; Wang, C.;
efficient oxidative dehydrogenation. J. Am. Chem. Soc. 2022, 144,
Huang, H.; Yang, C.; Ma, D.; Li, Q.; et al. Light-induced redox
16855−16865.
looping of a rhodium/CexWO3 photocatalyst for highly active and
(35) Wang, L.; Yan, T.; Song, R.; Sun, W.; Dong, Y.; Guo, J.; Zhang,
robust dry reforming of methane. Angew. Chem., Int. Ed. 2022, 61,
Z.; Wang, X.; Ozin, G. A. Room-temperature activation of H2 by a
No. e202200567.
surface frustrated Lewis pair. Angew. Chem., Int. Ed. 2019, 58, 9501−
(18) Lv, K.; Teng, C.; Shi, M.; Yuan, Y.; Zhu, Y.; Wang, J.; Kong, Z.;
9505.
Lu, X.; Zhu, Y. Hydrophobic and electronic properties of the E-MoS2 (36) Andrews, L.; Wang, X. Infrared spectra of indium hydrides in
nanosheets induced by FAS for the CO2 electroreduction to syngas solid hydrogen and of solid Indane. Angew. Chem., Int. Ed. 2004, 43,
with a wide range of CO/H2 ratios. Adv. Funct. Mater. 2018, 28, 1706.
1802339. (37) Maeno, Z.; Yasumura, S.; Wu, X.; Huang, M.; Liu, C.; Toyao,
(19) Xing, M.; Zhang, J.; Chen, F.; Tian, B. An economic method to T.; Shimizu, K.-I. Isolated indium hydrides in CHA zeolites:
prepare vacuum activated photocatalysts with high photo-activities speciation and catalysis for nonoxidative dehydrogenation of ethane.
and photosensitivities. Chem. Commun. 2011, 47, 4947−4949. J. Am. Chem. Soc. 2020, 142, 4820−4832.
(20) Wei, W.; Wei, Z.; Li, R.; Li, Z.; Shi, R.; Ouyang, S.; Qi, Y.; (38) Tan, T. H.; Xie, B.; Ng, Y. H.; Abdullah, S. F. B.; Tang, H. Y.
Philips, D. L.; Yuan, H. Subsurface oxygen defects electronically M.; Bedford, N.; Taylor, R. A.; Aguey-Zinsou, K.-F.; Amal, R.; Scott, J.
interacting with active sites on In2O3 for enhanced photo- Unlocking the potential of the formate pathway in the photo-assisted
thermocatalytic CO2 reduction. Nat. Commun. 2022, 13, 3199. Sabatier reaction. Nat. Catal. 2020, 3, 1034−1043.
(21) Jin, Z.; Wang, L.; Zuidema, E.; Mondal, K.; Zhang, M.; Zhang, (39) Graciani, J.; Mudiyanselage, K.; Xu, F.; Baber, A. E.; Evans, J.;
J.; Wang, C.; Meng, X.; Yang, H.; Mesters, C.; et al. Hydrophobic Senanayake, S. D.; Stacchiola, D. J.; Liu, P.; Hrbek, J.; Sanz, J. F.; et al.
zeolite modification for in situ peroxide formation in methane Highly active copper-ceria and copper-ceria-titania catalysts for
oxidation to methanol. Science 2020, 367, 193−197. methanol synthesis from CO2. Science 2014, 345, 546−550.
(22) Chen, L.; Huang, S.; Ras, R. H. A.; Tian, X. Omniphobic liquid- (40) Kattel, S.; Yan, B.; Yang, Y.; Chen, J. G.; Liu, P. Optimizing
like surfaces. Nat. Rev. Chem 2023, 7, 123−137. binding energies of key intermediates for CO2 hydrogenation to
(23) Marshall, S. T.; O’Brien, M.; Oetter, B.; Corpuz, A.; Richards, methanol over oxide-supported copper. J. Am. Chem. Soc. 2016, 138,
R. M.; Schwartz, D. K.; Medlin, J. W. Controlled selectivity for 12440−12450.
palladium catalysts using self-assembled monolayers. Nat. Mater. (41) Shen, C.; Sun, K.; Zhang, Z.; Rui, N.; Jia, X.; Mei, D.; Liu, C.
2010, 9, 853−858. Highly active Ir/In2O3 catalysts for selective hydrogenation of CO2 to
(24) Swedlund, P. J.; Miskelly, G. M.; McQuillan, A. J. Silicic acid methanol: experimental and theoretical studies. ACS Catal. 2021, 11,
adsorption and oligomerization at the ferrihydrite-water interface: 4036−4046.
interpretation of ATR-IR spectra based on a model surface structure. (42) Deng, B.; Song, H.; Wang, Q.; Hong, J.; Song, S.; Zhang, Y.;
Langmuir 2010, 26, 3394−3401. Peng, K.; Zhang, H.; Kako, T.; Ye, J. Highly efficient and stable
(25) Jiang, X.; Nie, X.; Gong, Y.; Moran, C. M.; Wang, J.; Zhu, J.; photothermal catalytic CO2 hydrogenation to methanol over Ru/
Chang, H.; Guo, X.; Walton, K. S.; Song, C. A combined experimental In2O3 under atmospheric pressure. Appl. Catal., B 2023, 327, 122471.
and DFT study of H2O effect on In2O3/ZrO2 catalyst for CO2 (43) Wang, L.; Dong, Y.; Yan, T.; Hu, Z.; Ali, F. M.; Meira, D. M.;
hydrogenation to methanol. J. Catal. 2020, 383, 283−296. Duchesne, P. N.; Loh, J. Y. Y.; Qiu, C.; Storey, E. E.; et al. Black

26059 https://doi.org/10.1021/jacs.3c07349
J. Am. Chem. Soc. 2023, 145, 26052−26060
Journal of the American Chemical Society pubs.acs.org/JACS Article

indium oxide a photothermal CO2 hydrogenation catalyst. Nat.


Commun. 2020, 11, 2432.

26060 https://doi.org/10.1021/jacs.3c07349
J. Am. Chem. Soc. 2023, 145, 26052−26060

You might also like