You are on page 1of 7

Research Article

Chemistry—A European Journal doi.org/10.1002/chem.202104514

www.chemeurj.org

MOF Encapsulated AuPt Bimetallic Nanoparticles for


Improved Plasmonic-induced Photothermal Catalysis of
CO2 Hydrogenation
Yaqin Wang,[a] Xibo Zhang,[a] Kuan Chang,[a] Zhiying Zhao,[a] Jiayu Huang,[a] and Qin Kuang*[a]

Abstract: Exploring new catalytic strategies for achieving ceeded those obtained by Au@Pt@UiO-66-NH2 with Pt shell
efficient CO2 hydrogenation under mild conditions is of great on Au (599 μmol gmetal 1 h 1) and Au@UiO-66-NH2
significance for environmental remediation. Herein, a compo- (218 μmol gmetal 1 h 1). The outstanding performances of AuP-
site photocatalyst Zr-based MOF encapsulated plasmonic t@UiO-66-NH2 were attributed to the synergetic effect
AuPt alloy nanoparticles (AuPt@UiO-66-NH2) was successfully originating from the plasmonic metal Au, doped active metal
constructed for the efficient photothermal catalysis of CO2 Pt, and encapsulation structure of UiO-66-NH2 shell. This work
hydrogenation. Under light irradiation at 150 °C, AuPt@UiO- provides a new way for photothermal catalysis of CO2 and a
66-NH2 achieved a CO production rate of reference for the design of high-performance plasmonic
1451 μmol gmetal 1 h 1 with 91 % selectivity, which far ex- catalysts.

Introduction photocatalysts.[8] Notably, such intermetallic effect strongly


depends on the geometry of bimetallic nanostructures (such as
CO2 is a well-known excess substance and chemical feedstock core-shell and alloy), which greatly affect the performance in
that could be converted into other value-added products.[1] For photocatalysis.[7,9] In this regard, the ideal bimetallic nano-
example, high purity CO issued from CO2 hydrogenation by structures should maintain sufficient plasmonic cross-sections
reverse water gas shift (RWGS) could be used as raw material in for light-harvesting while possessing a large number of active
Fischer-Tropsch synthesis.[2] However, efficient and highly catalytic sites for target reactions. In addition, for CO2 hydro-
selective CO2 hydrogenation under mild conditions is still genation, how to improve the adsorption capacity of CO2 on
challenging due to the high activation energy barrier.[3] In plasmonic metal catalysts is another problem to be solved.[10]
recent years, as an emerging route for CO2 hydrogenation, This is because the weak interaction between them often leads
photothermal catalysis has attracted more and more attention to low catalytic activity.
because of its unique advantages in reducing the temperature Herein, a unique composite photocatalyst based on metal–
required for the reaction by introducing light. Nevertheless, organic frameworks (MOFs) encapsulated plasmonic AuPt alloy
high-performance catalysts are still needed to be explored for NPs was constructed for the efficient CO2 hydrogenation under
the successful application of this route.[4] mild conditions. NH2-functionalized Zr-based MOFs (UiO-66-
Localized surface plasmon resonance (LSPR) of plasmonic NH2) were used to encapsulate AuPt alloy NPs to prevent their
metal (such as Au and Ag) nanoparticles (NPs) may benefit aggregation, as well as improve the adsorption of CO2.[11]
photocatalysis.[5] However, those plasmonic metals generally Strikingly, this well-designed plasmonic AuPt@UiO-66-NH2 cata-
exhibit low activity despite showing initial promise for plasmon- lyst achieved the CO production rate of 1451 μmol gmetal 1 h 1
induced photocatalytic chemistry.[6] In recent years, bimetallic under light irradiation at relatively low temperature (150 °C),
nanostructures that combine the optical behavior of plasmonic with 91 % selectivity. This activity of AuPt@UiO-66-NH2 is 2.4-
metal with reactive behavior of catalytic metal (such as Pt and fold and 6.6-fold higher than that of Au@Pt@UiO-66-NH2 and
Pd) attracted considerable interests due to their fantastic optical Au@UiO-66-NH2 at the same condition, respectively. The
properties and promising potential in photocatalysis.[7] Conse- intrinsic mechanism behind the outstanding performance of
quently, tremendous efforts have been devoted to the con- AuPt@UiO-66-NH2 was further clarified by studying the influen-
trolled construction of various bimetallic plasmonic ces of bimetallic structure difference and metal loading on the
reaction.
[a] Y. Wang, X. Zhang, K. Chang, Z. Zhao, J. Huang, Prof. Q. Kuang
State Key Laboratory of Physical Chemistry of Solid Surfaces, Collaborative
Innovation Center of Chemistry for Energy Materials, and Department of Results and Discussion
Chemistry
College of Chemistry and Chemical Engineering, Xiamen University
To investigate the influences of bimetallic structures on the
Xiamen 361005 (P. R. China)
E-mail: akiwang98@126.com catalytic performances, AuPt alloy NPs and Au@Pt core-shell
Supporting information for this article is available on the WWW under NPs were pre-prepared by co-reduction route and Au-seeded
https://doi.org/10.1002/chem.202104514 growth route, respectively (Figure 1a and 1b). Both obtained

Chem. Eur. J. 2022, 28, e202104514 (1 of 6) © 2022 Wiley-VCH GmbH


15213765, 2022, 16, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202104514 by UFPI - Universidade Federal do Piaui, Wiley Online Library on [27/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104514

Figure 1. The synthetic routes of (a) AuPt alloy NPs and (b) Au@Pt core-shell NPs. TEM and HRTEM images of (c, d) AuPt alloy NPs and (e, f) Au@Pt core-shell
NPs. Insets in (c, e) are corresponding particle size distributions, and insets in (d, f) are enlarged HRTEM images of the region marked with a square. HAADF-
STEM images and EDS elemental maps of (g) AuPt alloy NPs and (i) Au@Pt core-shell NPs. EDS line scanning profiles of selected single particles in (g) and (i),
respectively.

Au Pt NPs showed uniform sizes. However, the particle sizes of associated with the LSPR peak of Au. By comparison, AuPt alloy
Au@Pt core-shell NPs (15.9 nm) were slightly larger than those NPs revealed slightly red-shifted and broadened absorption
of AuPt alloy NPs (14.8 nm) due to the seeded growth peak. While Au NPs are coated with a Pt shell, the shift and
(Figure 1c and 1e). Note that the power X-ray diffraction (XRD) broadening of the absorption peak become more obvious.
patterns of both samples did not differ from that of pure Au, The as-prepared Au Pt NPs were encapsulated within UiO-
which might be due to the low Pt content (Figure S1).[12] 66-NH2 with assistance of PVP, and the final products were
However, significant differences were seen in high resolution measured by scanning electron microscope (SEM) and TEM to
transmission electron microscopy (HR-TEM) images in terms of explore the structures. As shown in Figure 2a–b, AuPt@UiO-66-
spacings of the lattice fringes in surface regions of both samples NH2 exhibited irregular particles of around 100 ~ 150 nm, with
(Figure 1d and 1f). Meanwhile, Pt was also detected by X-ray metallic NPs individually located at the center of the MOF
photoelectron spectroscopy (XPS, Figure S2) and inductively matrix with low contrast. Due to the low dosage of Au Pt NPs,
coupled plasma-mass spectrometry (ICP-MS, Table S1). To the self-nucleation process of UiO-66-NH2 occurred in reaction
confirm the difference in surface composition and structure of solution inevitably, which accounted for nearly one fifth in the
both samples, energy dispersive X-ray spectroscopy (EDS) products. Such one-to-one encapsulation structure was also
elemental mapping and line scanning were conducted. For clearly reflected in the HAADF-STEM image of AuPt@UiO-66-
AuPt alloy NPs (Figure 1g and 1 h), the signals of Pt looked NH2, as well as the corresponding EDS elemental mapping
uniformly distributed in the whole particle and completely images (Figure 2c). Similarly, Au@Pt NPs and Au NPs were
overlapped with that of Au. By contrast, the signals of Pt were encapsulated in MOF in the same form (Figure S4). The XRD
more enriched on the surfaces of Au@Pt core-shell NPs analyses revealed that the shell consists of NH2-functionalized
(Figure 1i). Specifically, it was observed that the signals of Pt Zr-based MOFs (UiO-66-NH2) (Figure 2d). Compared to UiO-
risen outside that of Au in the EDS line scanning profiles of 66-NH2 (Figure 2e), the XPS analyses further indicated the shift
selected single Au@Pt particles, showing the typical character- in Zr 3d peaks of AuPt@UiO-66-NH2 and Au@Pt@UiO-66-NH2 to
istic of core-shell structure (Figure 1j). On the other hand, the lower binding energies by 0.1 and 0.4 eV, respectively. Thus,
difference in surface structure of metal NPs would influence the strong interactions existed between the metallic cores (AuPt
light absorption properties. As shown in Figure S3, Au NPs and Au@Pt) and MOF shell.[13] Note that the LSPR peak
displayed a strong absorption peak around 524 nm, which is

Chem. Eur. J. 2022, 28, e202104514 (2 of 6) © 2022 Wiley-VCH GmbH


15213765, 2022, 16, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202104514 by UFPI - Universidade Federal do Piaui, Wiley Online Library on [27/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104514

Figure 2. (a) SEM and (b) TEM images of AuPt@UiO-66-NH2. (c) HAADF-STEM image and corresponding EDS elemental maps of AuPt@UiO-66-NH2. (d) XRD
patterns and (e) Zr 3d XPS spectra of AuPt@UiO-66-NH2, Au@Pt@UiO-66-NH2, and UiO-66-NH2. (f) solid UV-vis spectra of AuPt@UiO-66-NH2, Au@Pt@UiO-66-
NH2, Au@UiO-66-NH2, and UiO-66-NH2.

originating from Au NPs was well preserved after encapsulation increased to 1451 μmol gmeta 1 h 1 and 137 μmol gmetal 1 h 1,
within MOF, while showing shifts to ca. 550 nm (Figure 2f). respectively. The CO selectivity, in this case, reached 91 %.
The catalytic activities of the as-obtained catalysts toward Clearly, light excitation played an important role in the reaction.
CO2 hydrogenation were studied in vapor-solid reaction mode To verify the intrinsic role of light in photocatalysis, a wave-
using a commercial evaluation system. To this end, CO2 and H2 length-dependent catalytic reaction was also performed using
at the molar ratio of 1 : 3 were used as reactants. In a typical filters (λ < 400 nm and 400 < λ < 650 nm). The activity of
catalytic test, the reaction was carried out at 150 °C under AuPt@UiO-66-NH2 under visible light was about 4.3-fold higher
irradiation with a 300 W Xe lamp for 4 h. The performances of than that obtained under ultraviolet light, indicating the
the catalyst with encapsulated AuPt alloy NPs (i. e., AuPt@UiO- dominant promoting role of LSPR in the visible region from
66-NH2) were first investigated. As shown in Figure 3a, no AuPt alloy NPs on light-driven CO2 hydrogenation.
product was detected after 4 h reaction without irradiation at Previous studies showed a linear dependence of reaction
150 °C, indicating that it was difficult to activate CO2 using only rate on reaction temperature in plasmonic photocatalysts.[14]
heat at this temperature. Upon lighting, the yield of CO and CH4 Therefore, the effect of reaction temperature on the photo-
catalytic CO2 hydrogenation activities was further investigated.
As expected, the CO and CH4 production rates increased linearly
with temperature (Figure 3b). In addition, higher reaction
temperatures led to greater CO selectivity since RWGS reaction
(CO2 + 2H2 = CO + 2H2O) as an endothermic reaction more likely
occurred at higher temperatures. The stability of AuPt@UiO-66-
NH2 was also studied by time-dependent experiments. As
depicted in Figure S5, CO production rate over AuPt@UiO-66-
NH2 varied linearly with the reaction time. Besides, the XRD
patterns and TEM images of AuPt@UiO-66-NH2 before and after
4 h of reaction showed no significant changes (Figure S6-7).
Thus, AuPt@UiO-66-NH2 possessed high stability.
As shown in UV-vis absorption spectra (Figure 2e and
Figure 3. (a) The production rates of CO2 hydrogenation over AuPt@UiO-66-
NH2 at 150 °C under different light irradiations (UV: λ < 400 nm, Vis: 400 < Figure S3), LSPR of plasmonic metal NPs was sensitive to the
λ < 650 nm) for 4 h. (b) The correlation between the production rates and composition and structure. As a result, the catalytic perform-
temperature for AuPt@UiO-66-NH2. ances of the as-obtained metal@MOF composites toward CO2

Chem. Eur. J. 2022, 28, e202104514 (3 of 6) © 2022 Wiley-VCH GmbH


15213765, 2022, 16, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202104514 by UFPI - Universidade Federal do Piaui, Wiley Online Library on [27/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104514

hydrogenation were closely related to the plasmonic metal NPs


encapsulated within UiO-66-NH2. This was verified by testing
the catalysts encapsulated with other Au-based NPs, such as
Au@Pt@UiO-66-NH2 and Au@UiO-66-NH2. Note that the per-
formance of Au@Pt@UiO-66-NH2 encapsulated with Au@Pt
core-shell NPs (CO production rate = 599 μmol gmetal 1 h 1, CH4
production rate = 108 μmol gmetal 1 h 1) showed significant im-
provement when compared to Au@UiO-66-NH2 (CO production
rate = 218 μmol gmetal 1 h 1), but still far lower than AuPt@UiO-
66-NH2. Thus it can be concluded that the composition and
structure of plasmonic metal NPs were important in the
resulting catalytic activity, and AuPt alloy was a more favorable
form of plasmonic catalysts than Au@Pt core-shell structure.
It should be mentioned that UiO-66-NH2 matrix itself
possessed no activity toward CO2 hydrogenation. What is the
role of the UiO-66-NH2 matrix? Consequently, the influence of
the loading style of AuPt alloy NPs on hydrogenation activity
was also investigated. As illustrated in Figure 4b, AuPt NPs
supported on UiO-66-NH2 (AuPt/UiO-66-NH2) surface resulted in
CO production rate of only 244 μmol gmetal 1 h 1, with CO
selectivity dramatically decreased to 66 %. Thus, the encapsula-
tion structure of metal@MOF attributed to the catalytic
Scheme 1. Schematic illustration of the mechanism of plasmonic-induced
efficiency. Besides, when the support was changed to SiO2 and photothermal catalytic CO2 hydrogenation over AuPt@UiO-66-NH2.
ZrO2, the catalysts (i. e., AuPt/SiO2 and AuPt/ZrO2) showed
almost no catalytic activities. When combined together, the
above results demonstrated that the CO production was
essentially enhanced by the synergistic effect between UiO-66- hydrogenation reactions, but the catalytic performance can
NH2 and AuPt alloy NPs. greatly be improved after coupling with catalytically active Pt.[16]
To the best of our knowledge, the activity performance of In this work, the AuPt alloy structure was found superior to that
AuPt@UiO-66-NH2 was much better than that of most catalysts of Au@Pt core-shell. The Pt shell caused LSPR red-shifting in
reported recently (Table S2). The outstanding plasmonic-in- energy, and broadened the light absorption range. However, it
duced photothermal catalytic performance of AuPt@UiO-66-NH2 inhibited LSPR effect of Au NPs to some extent, thereby
toward CO2 hydrogenation was attributed to several factors, reducing the production efficiency of hot electrons. By contrast,
including the plasmonic property of Au, catalytic role of Pt, and the AuPt alloy structure introduced active metal into Au NPs
encapsulation function of MOF (Scheme 1). First, the chemical without affecting the LSPR ability, more conducive to promot-
transformation over AuPt@UiO-66-NH2 was mainly driven by ing the conversion efficiency of CO2. Third, the encapsulation
high-energy hot electrons generated by LSPR effect of Au structure in AuPt@UiO-66-NH2 catalyst played the role of reactor
NPs.[15] Compared to thermal catalysis, the introduction of Au to provide a unique micro-environment suitable for CO2
LSPR effect can greatly reduce the activation energy required hydrogenation.[17] On the one hand, the MOF shell significantly
for the reaction, so that CO2 conversion can be carried out improved the adsorption capability toward CO2 due to the
efficiently under mild conditions. Second, Au was inert during ultrahigh surface area and alkalinity of UiO-66-NH2.[18] On the
other hand, the highly porous structure also provided well-
confined space for the diffusion and collision of CO2 molecules
with plasmonic AuPt NPs.[19] Furthermore, such special confine-
ment effect controlled the adsorption orientation of reactants
and products on metal NPs surface, thereby significantly
improving the product selectivity.[20]

Conclusion

AuPt@UiO-66-NH2 composite catalyst was successfully synthe-


sized for efficient plasmonic-induced photothermal catalysis of
CO2 hydrogenation. Under light irradiation at 150 °C, the CO
Figure 4. (a) Performances of AuPt@UiO-66-NH2, Au@Pt@UiO-66-NH2,
Au@UiO-66-NH2, and UiO-66-NH2 toward hydrogenation of CO2 under light production rate over AuPt@UiO-66-NH2 reached
irradiation. (b) The production rates obtained by the prepared catalysts 1451 μmol gmetal 1 h 1, a value 2.4-fold and 6.6-fold higher than
consisting of different loading styles. that of Au@Pt@UiO-66-NH2 and Au@UiO-66-NH2, respectively.

Chem. Eur. J. 2022, 28, e202104514 (4 of 6) © 2022 Wiley-VCH GmbH


15213765, 2022, 16, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202104514 by UFPI - Universidade Federal do Piaui, Wiley Online Library on [27/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104514

The outstanding performances of AuPt@UiO-66-NH2 were on the efficiency of catalysis was evaluated using UV cut-off filter
attributed to the synergetic effect induced by plasmonic metal (λ < 400 nm) and (400 < λ < 650 nm) filters. After the completion of
Au, active metal Pt, and versatile substrate UiO-66-NH2. the reaction, the content of gas-phase was determined by gas
chromatography (GC, Shimadzu GC-2014) using a flame-ionized
Compared to Au@Pt core-shell NPs, AuPt alloy NPs possessed detector.
also a more ideal plasmonic catalyst structure for photothermal
catalytic CO2 hydrogenation, due to advantages in retaining
both the optical behavior of plasmonic metal and reactive
behavior of catalytic metal. In sum, better insights into the Acknowledgements
structure-performance relationship of plasmonic metal@MOF
composite photocatalysts were provided in this work, and the This work was supported by the National Key Research and
suggested construction strategy looks promising for wider Development Program of China (2017YFA0206801), and the
applicability in more emerging plasmonic photocatalytic sys- National Natural Science Foundation of China (No. 22071202,
tems. 21931009, 21721001, and 21773190).

Experimental Section Conflict of Interest


Preparation of Au NPs and AuPt alloy NPs. Firstly, 1 m L HAuCl4
aqueous solution (0.025 M) and 100 mL de-ionized water were The authors declare no conflict of interest.
poured into a 150 mL three-necked flask followed by heating to
100 °C. Subsequently, 3 mL citrate sodium aqueous solution
(0.034 M) was added to the solution in one shot. After heating for Data Availability Statement
another 20 min under vigorous stirring, deep-red citrate-stabilized
AuNPs were formed. Next, the resulting solution was cooled to
room temperature, then 500 mg PVP was added and left under Research data are not shared.
stirring for 24 h. The resulting Au NPs were collected by
centrifugation at 19000 rpm from the AuNPs colloidal solution Keywords: CO2 hydrogenation · localized surface plasmon
(100 mL) and then dispersed in DMF (1 mL). The purplish-red AuPt
resonance · metal-organic frameworks · photothermal catalysis
alloy NPs were obtained by using a mixture of 0.3 mL H2PtCl6
aqueous solution (0.019 M) and 1 mL HAuCl4 aqueous solution
(0.025 M).
[1] a) R. P. Ye, J. Ding, W. Gong, M. D. Argyle, Q. Zhong, Y. Wang, C. K.
Preparation of Au@Pt core-shell NPs. The as-obtained Au NPs Russell, Z. Xu, A. G. Russell, Q. Li, M. Fan, Y. G. Yao, Nat. Commun. 2019,
dispersion was further diluted by 20 mL DMF, heated to 100 °C 10, 5698; b) C. Hepburn, E. Adlen, J. Beddington, E. A. Carter, S. Fuss, N.
under stirring, and supplemented by 0.3 mL H2PtCl6 aqueous Mac Dowell, J. C. Minx, P. Smith, C. K. Williams, Nature 2019, 575, 87–97.
solution (0.019 M). After reaction for 1 h, the color of the solution [2] a) Y. A. Daza, J. N. Kuhn, RSC Adv. 2016, 6, 49675–49691; b) M. D.
changed from deep-red to deep-purple. The resulting Au@Pt NPs Porosoff, B. Yan, J. G. Chen, Energy Environ. Sci. 2016, 9, 62–73; c) A. M.
were then collected by centrifugation at 19000 rpm from the Bahmanpour, M. Signorile, O. Kröcher, Appl. Catal. B 2021, 295, 120319.
[3] W. Zhou, K. Cheng, J. Kang, C. Zhou, V. Subramanian, Q. Zhang, Y.
resulting colloidal solution (100 mL), followed by dispersion in 1 mL Wang, Chem. Soc. Rev. 2019, 48, 3193–3228.
DMF for later use. [4] a) C. Wang, Z. Sun, Y. Zheng, Y. H. Hu, J. Mater. Chem. A 2019, 7, 865–
887; b) Z. Yang, Y. Qi, F. Wang, Z. Han, Y. Jiang, H. Han, J. Liu, X. Zhang,
Preparation of AuPt@UiO-66-NH2, Au@Pt@UiO-66-NH2, and
W. J. Ong, J. Mater. Chem. A 2020, 8, 24868–24894; c) J. D. Yi, R. K. Xie,
Au@UiO-66-NH2. Firstly, 1 mL AuPt NPs dispersed in DMF, 27.3 mg Z. L. Xie, G. L. Chai, T. F. Liu, R. P. Chen, Y. B. Huang, R. Cao, Angew.
H2ATA, 100 mg PVP, and 1.2 mL acetic acid were added to 7 mL Chem. Int. Ed. 2020, 59, 23641–23648; Angew. Chem. 2020, 132, 23849–
DMF and stirred under ultrasounds for 15 min. Meanwhile, 33.4 mg 23856.
ZrCl4 was dispersed in 7 mL DMF under ultrasounds to yield a [5] a) S. Linic, P. Christopher, D. B. Ingram, Nat. Mater. 2011, 10, 911–921;
uniform dispersion. Next, the above two solutions were mixed in a b) U. Aslam, V. G. Rao, S. Chavez, S. Linic, Nat. Catal. 2018, 1, 656–665.
25 mL Teflon-lined stainless steel autoclave under heating to 120 °C [6] a) V. G. Rao, U. Aslam, S. Linic, J. Am. Chem. Soc. 2019, 141, 643–647;
for 24 h. After washing several times with DMF and drying under b) S. Shuang, R. Lv, Z. Xie, Z. Zhang, Sci. Rep. 2016, 6, 26670; c) W. Yang,
J. Zhao, H. Tian, L. Wang, X. Wang, S. Ye, J. Liu, J. Huang, Small 2020, 16,
vacuum at 120 °C for 10 h, the AuPt@UiO-66-NH2 was obtained. The 2002236.
synthetic procedures of Au@Pt@UiO-66-NH2 and Au@UiO-66-NH2 [7] K. Sytwu, M. Vadai, J. A. Dionne, Adv. Phys-X 2019, 4, 1619480.
were similar to that of Au@UiO-66-NH2, except that AuPt NPs were [8] a) H. Song, X. Meng, T. D. Dao, W. Zhou, H. Liu, L. Shi, H. Zhang, T.
replaced with Au@Pt or Au NPs. Nagao, T. Kako, J. Ye, ACS Appl. Mater. 2018, 10, 408–416; b) P. Liu, X.
Gu, H. Zhang, J. Cheng, J. Song, H. Su, Appl. Catal. B 2017, 204, 497–504;
Catalytic activity testing. The hydrogenation of CO2 was conducted c) J. Mo, E. C. M. Barbosa, S. Wu, Y. Li, Y. Sun, W. Xiang, T. Li, S. Pu, A.
in a vapor-solid reaction mode using a commercial evaluation Robertson, T. S. Wu, Y. l. Soo, T. V. Alves, P. H. C. Camargo, W. Kuo,
system (CEL-HPR100T + , Beijing China Education Au-Light Co., Ltd, S. C. E. Tsang, Adv. Funct. Mater. 2021, 31, 2102517; d) X. B. Zhang, Y. Y.
see Figure S8). In a typical reaction, 5 mg catalyst was first evenly Fan, E. M. You, Z. X. Li, Y. D. Dong, L. N. Chen, Y. Yang, Z. X. Xie, Q.
dispersed on a quartz chip and then transferred into a 150 mL Kuang, L. S. Zheng, Nano Energy 2021, 84, 105950.
stainless steel reactor. Afterward, the stainless steel was pressurized [9] D. Manchon, J. Lerme, T. Zhang, A. Mosset, C. Jamois, C. Bonnet, J. M.
Rye, A. Belarouci, M. Broyer, M. Pellarin, E. Cottancin, Nanoscale 2015, 7,
with CO2 (0.25 MPa) and H2 (0.75 MPa). The quartz window 1181–1192.
(diameter: 3.5 cm) at the top of stainless steel and compensatory [10] a) X. H. Liu, J. G. Ma, Z. Niu, G. M. Yang, P. Cheng, Angew. Chem. Int. Ed.
heating side allowed the passage of incident light from the 300 W Engl. 2014, 54, 988–991; Angew. Chem. 2014, 127, 1002–1005; b) H.
Xe lamp and extra heat into the reactor to participate in the Zhang, T. Wang, J. Wang, H. Liu, T. D. Dao, M. Li, G. Liu, X. Meng, K.
photothermal catalysis process. The influence of light wavelength Chang, L. Shi, T. Nagao, J. Ye, Adv. Mater. 2016, 28, 3703–3710; c) L. L.
Ling, W. Yang, P. Yan, M. Wang, H. L. Jiang, Angew. Chem. Int. Ed. 2022,

Chem. Eur. J. 2022, 28, e202104514 (5 of 6) © 2022 Wiley-VCH GmbH


15213765, 2022, 16, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/chem.202104514 by UFPI - Universidade Federal do Piaui, Wiley Online Library on [27/03/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Research Article
Chemistry—A European Journal doi.org/10.1002/chem.202104514

202116396; Angew. Chem. 2022, 202116396; d) D. X. Chen, W. J. Yang, L. Chang, S. Yao, W. Zhang, Y. Wang, P. Wang, W. B. Lin, T. B. Lu, Z. M.
Jiao, L. Y. Li, S. H. Yu, H. L. Jiang, Adv. Mater. 2020, 2000041. Zhang, J. Am. Chem. Soc. 2021, 143, 6114–6122; d) M. Liu, Y. F. Mu, S.
[11] a) W. Zhang, W. Shi, W. Ji, H. Wu, Z. Gu, P. Wang, X. Li, P. Qin, J. Zhang, Yao, S. Guo, X. W. Guo, Z. M. Zhang, T. B. Lu, Appl. Catal. B 2019, 245,
Y. Fan, T. Wu, Y. Fu, W. Zhang, F. Huo, ACS Catal. 2020, 10, 5805–5813; 496–501.
b) A. Zhou, Y. Dou, J. Zhou, J. R. Li, ChemSusChem 2020, 13, 205–211. [18] a) Q. Yang, W. Liu, B. Wang, W. Zhang, X. Zeng, C. Zhang, Y. Qin, X. Sun,
[12] D. Li, F. Meng, H. Wang, X. Jiang, Y. Zhu, Electrochim. Acta. 2016, 190, T. Wu, J. Liu, F. Huo, J. Lu, Nat. Commun. 2017, 8, 14429; b) F.
852–861. Vermoortele, B. Bueken, G. Le Bars, B. Van de Voorde, M. Vandichel, K.
[13] a) Q. Yang, Q. Xu, H. L. Jiang, Chem. Soc. Rev. 2017, 46, 4774–4808; b) H. Houthoofd, A. Vimont, M. Daturi, M. Waroquier, V. Van Speybroeck, C.
Liu, L. Chang, L. Chen, Y. Li, ChemCatChem 2016, 8, 946–951; c) C. He, J. Kirschhock, D. E. De Vos, J. Am. Chem. Soc. 2013, 135, 11465–11468.
Liang, Y. H. Zou, J. D. Yi, Y. B. Huang, R. Cao, Natl. Sci. Rev. 2021, [19] a) J. D. Xiao, Q. C. Shang, Y. J. Xiong, Q. Zhang, Y. Luo, S. H. Yu, H. L.
nwab157. Jiang, Angew. Chem. Int. Ed. 2016, 55, 9389–9393; Angew. Chem. 2016,
[14] a) P. Christopher, H. Xin, A. Marimuthu, S. Linic, Nat. Mater. 2012, 11, 128, 9535–9539; b) Y. Zhu, X. Qiu, S. Zhao, J. Guo, X. Zhang, W. Zhao, Y.
1044–1050; b) X. Zhang, X. Li, D. Zhang, N. Q. Su, W. Yang, H. O. Everitt, Shi, Z. Tang, Nano Res. 2020, 13, 1928–1932.
J. Liu, Nat. Commun. 2017, 8, 14542; c) Y. Li, M. Wen, Y. Wang, G. Tian, C. [20] a) B. Rungtaweevoranit, J. Baek, J. R. Araujo, B. S. Archanjo, K. M. Choi,
Wang, J. Zhao, Angew. Chem. Int. Ed. 2020, 60, 910–916; Angew. Chem. O. M. Yaghi, G. A. Somorjai, Nano Lett. 2016, 16, 7645–7649; b) Y. B.
2020, 133, 923–929. Huang, Q. Wang, J. Liang, X. S. Wang, R. Cao, J. Am. Chem. Soc. 2016,
[15] M. J. Kale, T. Avanesian, P. Christopher, ACS Catal. 2013, 4, 116–128. 138, 10104–10107; c) C. He, Q. J. Wu, M. J. Mao, Y. H. Zou, B. T. Liu, Y. B.
[16] a) D. Li, S. H. Yu, H. L. Jiang, Adv. Mater. 2018, 30, 1707377; b) W. Zhang, Huang, R. Cao, CCS Chem. 2020, 2, 2368–2380.
L. Wang, K. Wang, M. U. Khan, M. Wang, H. Li, J. Zeng, Small 2017, 13,
1602583.
[17] a) L. Y. Li, Z. X. Li, W. J. Yang, Y. M. Huang, G. Huang, Q. Q. Guan, Y. M.
Dong, J. L. Lu, S. H. Yu, H. L. Jiang, Chem 2021, 7, 686–698; b) S. S. Fu, S. Manuscript received: December 20, 2021
Yao, S. Guo, G. C. Guo, W. J. Yuan, T. B. Lu, Z. M. Zhang, J. Am. Chem. Accepted manuscript online: February 3, 2022
Soc. 2021, 143, 20792–20801; c) T. C. Zhuo, Y. Song, G. L. Zhuang, L. P. Version of record online: February 19, 2022

Chem. Eur. J. 2022, 28, e202104514 (6 of 6) © 2022 Wiley-VCH GmbH

You might also like