Issue |
Natl Sci Open
Volume 2, Number 2, 2023
Special Topic: Chemistry Boosts Carbon Neutrality
|
|
---|---|---|
Article Number | 20220064 | |
Number of page(s) | 14 | |
Section | Chemistry | |
DOI | https://doi.org/10.1360/nso/20220064 | |
Published online | 28 February 2023 |
RESEARCH ARTICLE
Hydroformylation over polyoxometalates supported single-atom Rh catalysts
1
Joint School of National University of Singapore and Tianjin University, International Campus of Tianjin University, Binhai New City, Fuzhou 350207, China
2
School of Materials Science and Engineering, Shaanxi Laboratory of Catalysis, Shaanxi University of Technology, Hanzhong 723001, China
3
School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, China
4
Department of Chemical & Biomolecular Engineering, National University of Singapore, Singapore 117580, Singapore
* Corresponding authors (emails: qiyu@snut.edu.cn (Qi Yu); ning.yan@nus.edu.sg (Ning Yan))
Received:
30
September
2022
Revised:
3
February
2023
Accepted:
6
February
2023
Atomic dispersion of Rh on phosphotungstic acid (PTA) salts was achieved by a self-assembled precipitation method using alkali metal ions as coprecipitation reagents. During styrene hydroformylation, the supported Rh single-atom catalyst (Rh1/M-PTA, M refers to an alkali metal ion) demonstrated an optimum turnover frequency (TOF) of 1076 h−1. With increasing ionic radius, the pore size of the catalysts increased in the following order: Rh1/K-PTA<Rh1/Rb-PTA<Rh1/Cs-PTA. The catalytic activity showed the same trend, suggesting a positive correlation between pore size and hydroformylation performance. Further experimental data suggested that temperature is an important factor affecting not only the activity but also the selectivity. This study enriches the understanding of the structure and catalytic properties of PTA-supported single-atom materials. The cation-controlled synthesis of catalysts may also be applied to prepare other single-atom catalysts with tunable pore size distributions.
Key words: polyoxometalates / Rh single-atom / hydroformylation / self-assembly / styrene
© The Author(s) 2023. Published by China Science Publishing & Media Ltd. and EDP Sciences.
This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
INTRODUCTION
The hydroformylation of olefins to aldehydes is among the largest industrial reactions promoted by homogeneous catalysts [1], enabling the functionalization of olefins and the growth of the carbon chain. It opens up paths towards a broad range of value-added downstream products such as amines, alcohols, esters, and carboxylic acids [2]. Despite their excellent activity and selectivity, traditional homogeneous rhodium catalysts are not easily recyclable and not sufficiently stable. Efforts have been made to immobilize rhodium complexes onto solid supports for easy separation, in which Rh species are still coordinated with organic ligands [3–6]. Alternatively, a phosphine ligand can be added in the postsynthesis stage to modulate Rh species on solid supports owing to its firm metal-ligand coordination [7]. After P doping, Rh exhibited excellent stabilization and enhanced regioselectivity through geometric and electronic effects [8,9].
Organic ligand-free single-atom catalysts (SACs) represent a new class of materials with multiple applications in catalysis [10–18]. Much progress has been made in unsaturated chemical hydrogenation [19–33], organic chemical oxidation [34–39], CO2 conversion [40–46], CO oxidation [47–56], biomass conversion [21,31,57–59], etc. Different supported Rh single-atom catalysts have also shown promising hydroformylation activity [60–65]. By depositing Rh onto the vacancies on ZnO nanowires (Rh1/ZnO-nw) [60] and CeO2 (Rh1/CeO2) [62,63] or replacing Co atoms with Rh on CoO nanosheets (Rh1/CoO) [61], the atomic dispersion of Rh was achieved with enhanced catalytic performance. Due to the strong metal-support interaction and unique local chemical environment, the Rh single-atom catalysts exhibited similar or even higher efficiency compared with the homogeneous Wilkinson’s catalyst and can also be reused with no significant loss of activity and selectivity. In addition to the metal oxide supports, a manganese-based metal organic framework (MOF)-incorporated Rh was designed and used for styrene hydroformylation at 80°C with high selectivity, high catalyst recyclability, and negligible metal leaching [66]. When ionic liquids (ILs) were added as extra protecting reagents, the Rh1/TiO2 catalyst demonstrated greatly enhanced stability without compromising activity [67] (turnover frequency, TOF = 878 h−1). After five runs, the TOF value was only slightly reduced by 9% to 800 h−1 during styrene hydroformylation. In a more recent study of the same reaction, Rh-substituted polyoxometalate-based ILs achieved ultrahigh activity (TOF>9000 h−1) [65]. Polyoxometalates (POMs) with quaternary ammonium cations showed a low melting point and behaved as homogeneous catalysts under reaction conditions.
Among various supports, POMs, as molecularly defined metal-oxo clusters, have been extensively used to prepare SACs with uniform structures. POMs only exhibit a limited number of binding sites for atomically dispersed metals such as Pt, Pd, or Rh, thus representing ideal systems for investigating the working mechanism of active sites on SACs. Previously, we studied CO and alcohol oxidation reactions [50,52,54,56], the hydrogenation of various unsaturated compounds [20], and the hydrogen spillover phenomenon and its correlation to catalysis [29]. Since the surface area of the parent molecule of POMs, usually referring to protonated heteropoly acids, is small and thus not suitable to be directly used as supports, two methods are used to increase their dispersion: (1) the anchoring of POMs on high surface area supports, such as carbon; (2) the formation of porous POM salts through ion exchange between protons and inorganic metal cations. The parent structure, comprising Keggin-type anions, formed new oriented and crystalline POM salts possessing larger pores and skeletons after the replacement of the original H5O2+ by cations (Cs+, , etc.), which to a large extent increased the specific surface area (from ~10 to ~100 m2/g) and pore size and therefore promoted the dispersion of metals. In this work, we prepared Rh1-doped, porous POM salts containing different cations and with pore sizes by a self-assembled precipitation method. Catalytic performance was evaluated under different conditions to probe structure-activity correlations.
EXPERIMENTAL
Catalyst preparation
The catalysts were synthesized by self-assembly through a typical procedure as follows. A specific amount of Rh(NO3)3 stock solution (Rh conc. = 1178.4 mg/L) was mixed with an aqueous solution of phosphotungstic acid (PTA) for 4 h at room temperature. Then, stoichiometric amounts of dissolved nitrates (CsNO3, RbNO3, KNO3) were added dropwise to the above mixture under vigorous stirring (600 r/min) for 20 min to form a white colloidal solution. After further stirring for 1 h, the generated precipitates were centrifuged (8000 r/min,10 min) and thoroughly washed to remove nitrate residues. After drying at 333 K overnight, the solids were ground, followed by calcination at 523 K for 30 min under flowing air. The obtained catalyst was denoted as xRh/M-PTA (where x represents weight percentage and M refers to the alkali metal cation Cs, Rb, or K).
Rh1/TiO2 Rh1/CeO2 and Rh1/ZrO2 with 0.1 wt% loadings of Rh were prepared by a simple impregnation method. In a typical procedure, stoichiometric amounts of Rh(acac)(CO)2 were dissolved in 10 mL of acetone, and 1 g of support (TiO2, CeO2, ZrO2) was added to the above Rh precursor solution. The obtained mixture was stirred in open air at room temperature until the solvent evaporated completely. Then, the obtained powder was calcined at 200°C for 1 h in air.
Catalyst characterization
The Rh contents in the catalyst samples were determined by inductively coupled plasma-mass spectrometry (ICP-MS) on an Agilent ICP-MS 7800 after microwave digestion. X-ray diffraction (XRD) patterns were obtained on a powder X-ray diffractometer (SmartLab, Rigaku) using Cu Kα radiation (λ = 1.5405 Å). Scanning electron microscopy (SEM) observations were conducted on a Hitachi Regulus 8100 instrument. The samples were first dispersed in ethanol and dropped onto a supported conductive adhesive. After solvent evaporation, the sample was treated by gold sputtering before SEM observation. Transmission electron microscopy (TEM) images were acquired using a JEOL JEM-F200 instrument. The samples were first dispersed in ethanol, after which a small amount of suspension was dropped onto carbon-coated copper grids followed by drying in air overnight. Element mapping was also conducted.
The specific surface areas and pore size distributions were estimated based on data acquired from a N2 adsorption analyser (BELSORP-max, MicrotracBEL). Prior to analysis, each sample was evacuated at 473 K for 2 h. N2 isotherms obtained at 77 K were analysed using the Horvath-Kawazoe (HK) method. CO-diffuse reflectance infrared Fourier transform (CO-DRIFT) spectra were acquired using an FTIR spectrometer (iS50 FT-IR, NICOLET) equipped with a liquid N2 cooled MCT detector and a DRIFT chamber (DRK-4-NI8, HARRICK). Prior to each analysis, the sample was pretreated under Ar flow (40 mL/min) at 473 K for 30 min and then cooled to room temperature. After supplying a CO/Ar mixture flow (6:60, v/v) to the cell for 20 min, spectra were recorded over 1 h while keeping the cell at 323 K to remove physically adsorbed CO.
X-ray adsorption spectra at the Rh K-edge were obtained on the BL01B1 beamline at the SPring-8 facility (Japan) operating in both fluorescence and transmittance modes. An electron beam energy of 8 GeV was used together with a Si (311) two-crystal monochromator.
Catalytic activity evaluation
The hydroxylation reaction was conducted in a sealed autoclave with a Teflon tube embedded inside. In a typical procedure, 20 mg catalyst was added to the tube, followed by introducing 2 mmol styrene and 3 mL toluene. After charging 2 MPa syngas (H2:CO = 1:1) into the autoclave, the sealed reactor was placed in the heater at the set temperature with vigorous stirring, and then the hydroformylation reaction started. At different time intervals, the reaction mixture was filtered prior to gas chromatography (GC) analysis using hexane as an internal standard.
Conversion % = converted substrates (mol) / total substrate (mol) × 100%.
Selectivity % = obtained products (mol) / total substrate (mol) × 100%.
TOF (h−1) = converted substrates (mol)/[total metal (mol) × reaction time (h)]; TOF was calculated based on low conversion (<15%). For better comparison, the specific reaction times were 2, 5 and 10 h for Rh1/Cs-PTA, Rh1/Rb-PTA, and Rh1/K-PTA, respectively.
Calculation details
The theoretical calculations were performed by using the spin-polarized Kahn-Sham formalism of density functional theory (DFT) via the Vienna ab initio simulation package (VASP) [68]. The Perdew-Burke-Ernzerhof (PBE) exchange-correlation functional of the generalized gradient approximation (GGA) was adopted [69]. Projector augmented-wave (PAW) potentials with scalar-relativistic effects were used to account for the interaction between the valence electrons and ionic core with the nucleus [70,71]. A cut-off energy of 420 eV was used for the plane wave expansion. The geometries were optimized with the self-consistent field energy converged to 10−5 eV, and the force convergence criteria were set to 0.02 eV/Å.
RESULTS AND DISCUSSION
Structure and texture of the catalyst
In the preparation process, the Rh3+ and phosphotungstate anions (PW12O403−) formed a neutral complex (Rh1PTA1) in a one-to-one manner [53]. By precipitating PW12O403− in solution with K+, Rb+ and Cs+, self-assembled Rh/M-PTA (M refers to Cs, Rb, or K) particles were subsequently obtained (Scheme S1). As shown in the SEM image, the particle size of the catalyst was distributed in the range of 100–200 nm (Figure 1A). Misono et al. [72] proposed that spherical self-assembled solids had three classes of structures, consisting of phosphotungstate anions (primary structure), ionic crystals (secondary structure), and assembled spherical porous particles (tertiary structure). In this study, rough surfaces with small aggregates were clearly observed because different crystalline grains tended to aggregate to form roughly round particles due to molecular cohesion in the aqueous solution (tertiary structure, Figure 1B). The ionic crystals had a cubic structure consisting of monovalent metal cations (K+, Rb+, Cs+) and PW12O403− after the ion exchange of protons (secondary structure, Figure 1C). In our previous work, based on extended X-ray absorption fine structure (EXAFS) data and DFT calculations, the Pt atoms [20,50] and Rh atoms [52] were most stable when located at 4-fold hollow sites on the POMs due to firm coordination. EXAFS spectra for the as-prepared samples were also provided in the present work. As shown in Figure S1, the intensity of Rh/Cs-PTA was similar to that of Rh2O3. According to the XANES, no Rh–Rh bond contribution was observed, indicating that Rh only existed in isolated Rh(III) species coordinated with O. The Rh–O shell fitting results are listed in Table S1. The Rh–O bond length and coordination number were calculated to be 2.02 Å and an average of 5.6, respectively.
Figure 1 (A) SEM image of Rh1/Cs-PTA. (B) Tertiary structure: porous particle (100–200 nm). (C) Secondary structure: ionic crystals (10–20 nm). (D) Primary structure: top view of the optimized geometry of Rh1/PTA, ~1 nm), purple: Rh, blue: W, red: O, pink: P. (E) Charge density difference of Rh1/PTA. (F) Spin-polarized partial density of states projected on Rh s, p and d (green, blue and red) states. |
The optimized geometry of Rh1/PTA is shown in Figure 1D. For the surface atoms of the Keggin-type structure, the bridges (Ob) and sides of the Keggin structure (Oc) were suitable anchoring sites for an isolated single metal atom. The Rh atom preferred to bind at the 4-fold hollow site of the PTA cluster, with a calculated binding energy of −2.48 eV. The optimized W–Oc and W–Ob bond lengths were 1.98 and 1.97 Å, respectively, indicating significant binding of a Rh single atom on the surface. The calculated bond angle of W–Oc–W (123°) was smaller than that of W–Ob–W (141°). The optimized Co1/PTA and Ir1/PTA catalysts are also shown in Figure S2. To gain more insight into the nature of the chemical bonding of Rh1/PTA, the charge density difference (CDD) was calculated. As shown in Figure 1E, there was a clear loss of electron density from the Rh atom, which implied that the electron density flowed from Rh to the nearest O atoms in Rh1/PTA. This charge transfer from the Rh centre to the Keggin-type PTA cluster cage caused ionic interactions between Rh and O atoms, which mainly accounted for the large binding energy of Rh [73]. The Bader charge of Rh1 in M1/PTA is +1.17|e|. We further studied the spin-polarized partial density of states (PDOS), as shown in Figure 1F. The spin-up and spin-down PDOS of the Rh 4d orbitals were asymmetric near the Fermi level, indicating the spin polarization of Rh. The presence of Rh 4d-based quantum states near the Fermi level resulted in high activity towards small molecules and may have played a role in activating the adsorbates during the catalytic reaction.
According to the N2 adsorption-desorption results, the catalyst had micropores ranging from 0.5 to 2 nm (Figure 2). The peak centred at 0.56 nm was attributed to the ordered pores in crystals, which depended on the size of the cations. With increasing ion size (K+<Rb+<Cs+), the micropores grew larger (Scheme S2). Additionally, there was a micropore distribution ranging from 0.6 to 2 nm, which was ascribed to the irregular slits between different crystals within the spherical particle (Figure 1C). In addition, the specific areas were calculated based on the N2 adsorption-desorption results. When protons were replaced by K+, the specific area increased from <10 to 147 m2/g. With increasing cation size, the specific areas of Rh1/Rb-PTA and Rh1/Cs-PTA were determined to be 163 and 160 m2/g, respectively.
Figure 2 (Upper) HK-plot analysis of N2 adsorption-desorption isotherms. (Lower) XRD patterns of Rh1/H-PTA (a), Rh1/K-PTA (b), Rh1/Rb-PTA (c), and Rh1/Cs-PTA (d). |
XRD patterns were recorded and are shown in Figure 2. The structure of the Keggin anion was well preserved in all samples. There was a significant shift of the crystalline peak to a small diffraction angle with increasing cation size. Brown et al. [74] confirmed the hexahydrate formula of (H5O2+)3(PW12O403−) through X-ray and neutron single-crystal diffraction methods. The six water molecules were paired in nearly planar diaquahydrogen ions of H5O2+. After the hydrated proton (H5O2+) was replaced by K+, Rb+, or Cs+, the lattice constants of Rh1/K-PTA, Rh1/Rb-PTA and Rh1/Cs-PTA decreased to 11.59, 11.68, and 11.92 Å, respectively. As a reference, the lattice constant was 12.16 Å when the hydrated proton (H5O2+) acted as the counter cation.
The element mapping image indicated that W, Cs, and Rh had high spatial correlation, demonstrating the uniform Cs-PTA structure and evenly distributed Rh (Figures 3A–3C). In the CO-DRIFT spectra (Figure 3D), two IR peaks at approximately 2110 and 2050 cm−1 were associated with the symmetric and asymmetric stretches of the positively charged Rh(CO)2 gem-dicarbonyl species [75], demonstrating that Rh metals were atomically dispersed on the supports.
Figure 3 (A–C) EDX mapping images of Rh1/Cs-PTA. (D) CO DRIFT spectra of different catalysts: Rh/K-PTA, Rh/Rb-PTA, Rh/Cs-PTA, and Rh/Cs-PTA. |
Hydroformylation activities
Due to the different solubilities of PTA salts (K-PTA>Rb-PTA>Cs-PTA) [72], fine particles were formed at different precipitation rates after the addition of cations into the PTA solution. As mentioned in the catalyst preparation section, Rh(NO3)3 was first mixed with an aqueous solution of PTA to form the Rh1-PTA1 complex through electrostatic interactions. Then, after the addition of cations (Cs+, Rb+, K+), the cation-PTA grains quickly precipitated due to their low solubility. During this self-assembly process, the initially formed Rh1-PTA1 complex could probably be coprecipitated and finally enclosed by the numerous crystal grains and aggregates of cation-PTA. According to the ICP-MS results, the Rh loading was significantly affected during the self-assembly process, which increased with the increasing precipitation rate of POM salt. The loading amount of Rh for Cs-PTA was 0.062%, higher than that for Rb-PTA (0.048%) and K-PTA (0.024%).
The hydroformylation of styrene was then conducted under different conditions (Table 1). Given that Rh loading was different for the three catalysts, TOF was calculated with respect to converted substrates per Rh site per unit time to evaluate the catalyst efficiency. Rh1/Cs-PTA exhibited the highest TOF (1076 h−1), followed by Rh1/Rb-PTA (581 h−1) and Rh1/K-PTA (202 h−1), under 2 MPa H2/CO (1:1) at 100°C (Table S2). Thus, TOF was cation dependent. As mentioned above, the micropore size was determined by monovalent cations with different radii (Table S4). For a fast reaction such as hydroformylation, it is not unreasonable to propose that the internal mass transfer of substrates, controlled by the pore size of the catalysts, plays a role in determining the catalyst activity. In the presence of Rh1/Cs-PTA with relatively larger pores, the mass transfer resistance was lower than that for Rh1/Rb-PTA and Rh1/K-PTA, resulting in higher catalytic efficiency. When allylbenzene and phenylbutene of a larger molecular size were applied as substrates, the activity obviously decreased since the allylbenzene and phenylbutene molecules had a higher resistance of mass transfer than styrene. The reusability of the catalyst was also evaluated. As shown in Figure S3, the catalyst lost its activity completely in the second run. The XRD results (Figure S4) confirmed that the Keggin structure was maintained after the reaction. However, the CO-DRIFT spectra indicated that there was no Rh remaining on the surface of the used catalyst (Figure S5). Moreover, the ICP results also confirmed that Rh was leached under hydroformylation conditions (Table S3). The hydroformylation of styrene over other metal oxide-supported catalysts was explored. CeO2- and TiO2-supported Rh single atoms have been investigated in previous papers (Table S5) and exhibited remarkable activity in this work (Table 1), while ZrO2-supported Rh catalysts were found to be less active. The catalytic performance of Rh single-atom catalysts is highly dependent on the support, which plays an important role between stability and activity [62].
Hydroformylation of styrene a
The effect of temperature was further investigated. In the presence of pressured syngas, CO molecules predominantly adsorbed on the Rh site over H2 due to stronger interactions at low temperature (80°C) [65]. After CO insertion, the hydrogenation of the carbonyl complex led to linear and/or branched aldehydes. As the temperature increased to 120°C, the activity increased, while the hydroformylation selectivity declined due to the increasing yield of the byproduct ethylbenzene. The hydrogenation of styrene was favoured probably because CO adsorption on Rh sites was weakened at elevated temperatures. Moreover, n-/iso-aldehyde was also favoured at 120°C. This might be because higher temperatures promoted the partial dissociation of the branched alkyl-rhodium intermediate, therefore preferentially enhancing k2reverse over k1reverse (Scheme 1), which was beneficial for the formation of n-aldehyde. The apparent activation energies of n-aldehyde and iso-aldehyde were further investigated. The results are shown in Figure S6. According to the exponential law of reaction rate:
Scheme 1 Reaction route of the hydroformylation of styrene and other substrates. |
lnTOF = −Ea / RT + C.
The activation energies of iso-aldehyde and n-aldehyde were 50.4 and 74.4 kJ/mol, respectively. The higher activation energy indicated that the production of iso-aldehyde is more favoured at lower temperatures, while n-aldehyde is preferably produced at higher temperatures. Therefore, the ratio of n-aldehyde to iso-aldehyde (n-/iso-) is temperature dependent. This calculation of apparent activation energy further supports the reaction routes in Scheme 1.
Based on the discussion above, a possible reaction mechanism is proposed (Scheme 1). Initially, the substrates passed through the outer surface to migrate to the enclosed Rh sites via the microslits in the aggregates and micropores present in crystals. On the Rh sites, the adsorption energy of styrene on Rh1/PTA was −4.21 eV (Figure S7). In the presence of CO, an alkyl-rhodium-carbonyl intermediate was first formed through either linear or branched coordination. After the additional CO insertion and sequential reduction by H-species, aldehyde was produced and desorbed to allocate coordination vacancies for the next styrene molecule. The major side reaction was the hydrogenation of the C=C double bond, which was favoured at high temperatures.
The leaching of Rh atoms in the POMs-supported catalyst is still a major challenge in this work. It is important to stabilize the Rh atoms without compromising activity during the practical liquid-phase hydroformylation reaction. Our previous work revealed that ILs are promising stabilizer to protect metal nanoparticles and SACs catalysts through electrostatic interaction [67,76]. For the self-assembled POMs-supported Rh catalysts, the combination of POMs and functional ILs will be investigated in the future work. The design of multi-functional ILs will be of great significance.
CONCLUSION
In this study, a porous POM salt-supported Rh single-atom catalyst was prepared to promote the hydroformylation reaction. The catalyst had three classes of structures consisting of PTA anions (primary structure), ionic crystals (secondary structure), and assembled spherical porous particles (tertiary structure). The Rh atoms were firmly coordinated onto the PTA surface within the crystals. During styrene hydroformylation, the countercation of PTA played a key role in regulating the pore size of the catalysts, which had a substantial influence on the catalytic activity. Rh1/Cs-PTA exhibited the highest activity (TOF = 1076 h−1) compared with Rh1/Rb-PTA and Rh1/K-PTA, while the stereoselectivity was not altered significantly (n-/iso-aldehyde = 1.2). The ratio of n-aldehyde to iso-aldehyde (n-/iso-) was found to be temperature dependent. The activation energies of iso-aldehyde and n-aldehyde were 50.4 and 74.4 kJ/mol, respectively. When the temperature was increased from 80 to 100°C, the chemoselectivity of aldehydes decreased from 98.8% to 88.7%. In contrast, the stereoselectivity of the linear products was favoured and increased from 1.2 to 1.4.
Acknowledgments
We gratefully acknowledge HZWTECH for providing computation facilities. The XAS spectra were acquired at the beam-line BL01B1 at SPring-8 (Japan Synchrotron Radiation Research Institute, Hyogo, Japan) under the approval of JASRI (Proposal No. 2022B1917). We also appreciate Sikai Wang for his professional operations and analysis in conducting the EXAFS experiments.
Funding
This work was supported by the National Natural Science Foundation of China (92061109 and 22273053) and the Natural Science Basic Research Program of Shaanxi (2021JCW-20 and 2022KJXX-18).
Author contributions
N.Y. conceived the idea. N.Y. and S.F. designed the research. S.F. did the experiments. Q.Y. and X.Y. carried out the theoretical calculations. S.F. and Q.Y. co-wrote the manuscript. X.M. revised the manuscript.
Conflict of interest
The authors declare no conflict of interest.
Supplementary information
The supporting information is available online at https://doi.org/10.1360/nso/20220064. The supporting materials are published as submitted, without typesetting or editing. The responsibility for scientific accuracy and content remains entirely with the authors.
References
- Franke R, Selent D, Börner A. Applied hydroformylation. Chem Rev 2012; 112: 5675-5732. [Article] [CrossRef] [PubMed] [Google Scholar]
- Börner A, Franke R. Hydroformylation: Fundamentals, Processes, and Applications in Organic Synthesis. John Weinheim: Wiley & Sons, 2016 [CrossRef] [Google Scholar]
- Li C, Wang W, Yan L, et al. A mini review on strategies for heterogenization of rhodium-based hydroformylation catalysts. Front Chem Sci Eng 2018; 12: 113-123. [Article] [CrossRef] [Google Scholar]
- Hanf S, Alvarado Rupflin L, Gläser R, et al. Current state of the art of the solid Rh-based catalyzed hydroformylation of short-chain olefins. Catalysts 2020; 10: 510. [Article] [CrossRef] [Google Scholar]
- Liu B, Wang Y, Huang N, et al. Heterogeneous hydroformylation of alkenes by Rh-based catalysts. Chem 2022; 8: 2630-2658. [Article] [CrossRef] [Google Scholar]
- Zhao K, Wang X, He D, et al. Recent development towards alkene hydroformylation catalysts integrating traditional homo- and heterogeneous catalysis. Catal Sci Technol 2022; 12: 4962-4982. [Article] [Google Scholar]
- Gao P, Liang G, Ru T, et al. Phosphorus coordinated Rh single-atom sites on nanodiamond as highly regioselective catalyst for hydroformylation of olefins. Nat Commun 2021; 12: 4698. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Liu B, Huang N, Wang Y, et al. Promotion of inorganic phosphorus on Rh catalysts in styrene hydroformylation: Geometric and electronic effects. ACS Catal 2021; 11: 1787-1796. [Article] [CrossRef] [Google Scholar]
- Huang N, Liu B, Lan X, et al. Promotion of diphosphine ligands (PPh2(CH2)PPh2, n = 1, 2, 3, 5, 6) for supported Rh/SiO2 catalysts in heterogeneous ethene hydroformylation. Mol Catal 2021; 511: 111736. [Article] [CrossRef] [Google Scholar]
- Yang XF, Wang A, Qiao B, et al. Single-atom catalysts: A new frontier in heterogeneous catalysis. Acc Chem Res 2013; 46: 1740-1748. [Article] [CrossRef] [PubMed] [Google Scholar]
- Wang A, Li J, Zhang T. Heterogeneous single-atom catalysis. Nat Rev Chem 2018; 2: 65-81. [Article] [CrossRef] [Google Scholar]
- Hülsey MJ, Zhang J, Yan N. Harnessing the wisdom in colloidal chemistry to make stable single-atom catalysts. Adv Mater 2018; 30: 1802304. [Article] [CrossRef] [Google Scholar]
- Lang R, Du X, Huang Y, et al. Single-atom catalysts based on the metal-oxide interaction. Chem Rev 2020; 120: 11986-12043. [Article] [Google Scholar]
- Li L, Chang X, Lin X, et al. Theoretical insights into single-atom catalysts. Chem Soc Rev 2020; 49: 8156-8178. [Article] [CrossRef] [PubMed] [Google Scholar]
- Zhang Q, Guan J. Applications of single-atom catalysts. Nano Res 2022; 15: 38-70. [Article] [NASA ADS] [CrossRef] [Google Scholar]
- Liu J. Catalysis by supported single metal atoms. ACS Catal 2017; 7: 34-59. [Article] [CrossRef] [Google Scholar]
- Ding S, Hülsey MJ, Pérez-Ramírez J, et al. Transforming energy with single-atom catalysts. Joule 2019; 3: 2897-2929. [Article] [CrossRef] [Google Scholar]
- Kaiser SK, Chen Z, Faust Akl D, et al. Single-atom catalysts across the periodic table. Chem Rev 2020; 120: 11703-11809. [Article] [CrossRef] [Google Scholar]
- Wei H, Liu X, Wang A, et al. FeOx-supported platinum single-atom and pseudo-single-atom catalysts for chemoselective hydrogenation of functionalized nitroarenes. Nat Commun 2014; 5: 5634. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Zhang B, Asakura H, Zhang J, et al. Stabilizing a platinum1 single-atom catalyst on supported phosphomolybdic acid without compromising hydrogenation activity. Angew Chem 2016; 128: 8459-8463. [Article] [NASA ADS] [CrossRef] [Google Scholar]
- Liu W, Chen Y, Qi H, et al. A durable nickel single-atom catalyst for hydrogenation reactions and cellulose valorization under harsh conditions. Angew Chem 2018; 130: 7189-7193. [Article] [NASA ADS] [CrossRef] [Google Scholar]
- He X, He Q, Deng Y, et al. A versatile route to fabricate single atom catalysts with high chemoselectivity and regioselectivity in hydrogenation. Nat Commun 2019; 10: 3663. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Zhang B, Sun G, Ding S, et al. Atomically dispersed Pt1-polyoxometalate catalysts: How does metal-support interaction affect stability and hydrogenation activity?. J Am Chem Soc 2019; 141: 8185-8197. [Article] [CrossRef] [PubMed] [Google Scholar]
- Zhang L, Zhou M, Wang A, et al. Selective hydrogenation over supported metal catalysts: From nanoparticles to single atoms. Chem Rev 2019; 120: 683-733. [Article] [Google Scholar]
- Long X, Li Z, Gao G, et al. Graphitic phosphorus coordinated single Fe atoms for hydrogenative transformations. Nat Commun 2020; 11: 4074. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Yang F, Ding S, Song H, et al. Single-atom Pd dispersed on nanoscale anatase TiO2 for the selective hydrogenation of phenylacetylene. Sci China Mater 2020; 63: 982-992. [Article] [CrossRef] [Google Scholar]
- Ma Y, Zhang X, Cao L, et al. Effects of the morphology and heteroatom doping of CeO2 support on the hydrogenation activity of Pt single-atoms. Catal Sci Technol 2021; 11: 2844-2851. [Article] [CrossRef] [Google Scholar]
- Wang L, Zhu C, Xu M, et al. Boosting activity and stability of metal single-atom catalysts via regulation of coordination number and local composition. J Am Chem Soc 2021; 143: 18854-18858. [Article] [CrossRef] [PubMed] [Google Scholar]
- Hülsey MJ, Fung V, Hou X, et al. Hydrogen spillover and its relation to hydrogenation: Observations on structurally defined single-atom sites. Angew Chem Int Ed 2022; 61: e202208237. [Article] [Google Scholar]
- Hung SF, Xu A, Wang X, et al. A metal-supported single-atom catalytic site enables carbon dioxide hydrogenation. Nat Commun 2022; 13: 819. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Ji K, Xu M, Xu SM, et al. Electrocatalytic hydrogenation of 5-hydroxymethylfurfural promoted by a Ru1Cu single-atom alloy catalyst. Angew Chem Int Ed 2022; 61: e202209849. [Article] [Google Scholar]
- Zhao E, Li M, Xu B, et al. Transfer hydrogenation with a carbon-nitride-supported palladium single-atom photocatalyst and water as a proton source. Angew Chem Int Ed 2022; 61: e202207410. [Article] [Google Scholar]
- Zhao L, Qin X, Zhang X, et al. A magnetically separable Pd single-atom catalyst for efficient selective hydrogenation of phenylacetylene. Adv Mater 2022; 34: 2110455. [Article] [NASA ADS] [CrossRef] [Google Scholar]
- Xie J, Yin K, Serov A, et al. Selective aerobic oxidation of alcohols over atomically-dispersed non-precious metal catalysts. ChemSusChem 2017; 10: 359-362. [Article] [CrossRef] [PubMed] [Google Scholar]
- Li T, Liu F, Tang Y, et al. Maximizing the number of interfacial sites in single-atom catalysts for the highly selective, solvent-free oxidation of primary alcohols. Angew Chem Int Ed 2018; 57: 7795-7799. [Article] [CrossRef] [Google Scholar]
- Luo L, Luo J, Li H, et al. Water enables mild oxidation of methane to methanol on gold single-atom catalysts. Nat Commun 2021; 12: 1218. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Shang Y, Xu X, Gao B, et al. Single-atom catalysis in advanced oxidation processes for environmental remediation. Chem Soc Rev 2021; 50: 5281-5322. [Article] [CrossRef] [PubMed] [Google Scholar]
- Wang H, Jiao L, Zheng L, et al. PdBi single-atom alloy aerogels for efficient ethanol oxidation. Adv Funct Mater 2021; 31: 2103465. [Article] [CrossRef] [Google Scholar]
- Zhang Z, Liu J, Wang J, et al. Single-atom catalyst for high-performance methanol oxidation. Nat Commun 2021; 12: 5235. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Su X, Yang XF, Huang Y, et al. Single-atom catalysis toward efficient CO2 conversion to CO and formate products. Acc Chem Res 2018; 52: 656-664. [Article] [Google Scholar]
- Gong L, Zhang D, Lin CY, et al. Catalytic mechanisms and design principles for single-atom catalysts in highly efficient CO2 conversion. Adv Energy Mater 2019; 9: 1902625. [Article] [CrossRef] [Google Scholar]
- Millet MM, Algara-Siller G, Wrabetz S, et al. Ni single atom catalysts for CO2 activation. J Am Chem Soc 2019; 141: 2451-2461. [Article] [CrossRef] [PubMed] [Google Scholar]
- Zhao C, Wang Y, Li Z, et al. Solid-diffusion synthesis of single-atom catalysts directly from bulk metal for efficient CO2 reduction. Joule 2019; 3: 584-594. [Article] [Google Scholar]
- Li M, Wang H, Luo W, et al. Heterogeneous single-atom catalysts for electrochemical CO2 reduction reaction. Adv Mater 2020; 32: 2001848. [Article] [CrossRef] [Google Scholar]
- Vijay S, Ju W, Brückner S, et al. Unified mechanistic understanding of CO2 reduction to CO on transition metal and single atom catalysts. Nat Catal 2021; 4: 1024-1031. [Article] [CrossRef] [Google Scholar]
- Zheng T, Liu C, Guo C, et al. Copper-catalysed exclusive CO2 to pure formic acid conversion via single-atom alloying. Nat Nanotechnol 2021; 16: 1386-1393. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Qiao B, Wang A, Yang X, et al. Single-atom catalysis of CO oxidation using Pt1/FeOx. Nat Chem 2011; 3: 634-641. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Qiao B, Liang JX, Wang A, et al. Single atom gold catalysts for low-temperature CO oxidation. Chin J Catal 2016; 37: 1580-1586. [Article] [CrossRef] [Google Scholar]
- Nie L, Mei D, Xiong H, et al. Activation of surface lattice oxygen in single-atom Pt/CeO2 for low-temperature CO oxidation. Science 2017; 358: 1419-1423. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Zhang B, Asakura H, Yan N. Atomically dispersed rhodium on self-assembled phosphotungstic acid: Structural features and catalytic CO oxidation properties. Ind Eng Chem Res 2017; 56: 3578-3587. [Article] [CrossRef] [Google Scholar]
- Zhang Z, Zhu Y, Asakura H, et al. Thermally stable single atom Pt/m-Al2O3 for selective hydrogenation and CO oxidation. Nat Commun 2017; 8: 16100. [Article] [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Hülsey MJ, Zhang B, Ma Z, et al. In situ spectroscopy-guided engineering of rhodium single-atom catalysts for CO oxidation. Nat Commun 2019; 10: 1330. [Article] [CrossRef] [PubMed] [Google Scholar]
- Hülsey MJ, Sun G, Sautet P, et al. Observing single-atom catalytic sites during reactions with electrospray ionization mass spectrometry. Angew Chem Int Ed 2021; 60: 4764-4773. [Article] [CrossRef] [PubMed] [Google Scholar]
- Hülsey MJ, Sun G, Sautet P, et al. Observing single-atom catalytic sites during reactions with electrospray ionization mass spectrometry. Angew Chem 2021; 133: 4814-4823. [Article] [CrossRef] [Google Scholar]
- Muravev V, Spezzati G, Su YQ, et al. Interface dynamics of Pd-CeO2 single-atom catalysts during CO oxidation. Nat Catal 2021; 4: 469-478. [Article] [CrossRef] [Google Scholar]
- Hülsey MJ, Baskaran S, Ding S, et al. Identifying key descriptors for the single-atom catalyzed CO oxidation. CCS Chem 2022; 4: 3296-3308. [Article] [CrossRef] [Google Scholar]
- Lu Y, Zhang Z, Wang H, et al. Toward efficient single-atom catalysts for renewable fuels and chemicals production from biomass and CO2. Appl Catal B-Environ 2021; 292: 120162. [Article] [CrossRef] [Google Scholar]
- De S, Burange AS, Luque R. Conversion of biomass-derived feedstocks into value-added chemicals over single-atom catalysts. Green Chem 2022; 24: 2267-2286. [Article] [CrossRef] [Google Scholar]
- Ge R, Wang Y, Li Z, et al. Selective electrooxidation of biomass-derived alcohols to aldehydes in a neutral medium: Promoted water dissociation over a nickel-oxide-supported ruthenium single-atom catalyst. Angew Chem Int Ed 2022; 61: e202200211. [Article] [Google Scholar]
- Lang R, Li T, Matsumura D, et al. Hydroformylation of olefins by a rhodium single-atom catalyst with activity comparable to RhCl(PPh3)3. Angew Chem Int Ed 2016; 55: 16054-16058. [Article] [CrossRef] [Google Scholar]
- Wang L, Zhang W, Wang S, et al. Atomic-level insights in optimizing reaction paths for hydroformylation reaction over Rh/CoO single-atom catalyst. Nat Commun 2016; 7: 14036. [Article] [Google Scholar]
- Amsler J, Sarma BB, Agostini G, et al. Prospects of heterogeneous hydroformylation with supported single atom catalysts. J Am Chem Soc 2020; 142: 5087-5096. [Article] [CrossRef] [PubMed] [Google Scholar]
- Li T, Chen F, Lang R, et al. Styrene hydroformylation with in situ hydrogen: Regioselectivity control by coupling with the low-temperature water-gas shift reaction. Angew Chem Int Ed 2020; 59: 7430-7434. [Article] [CrossRef] [PubMed] [Google Scholar]
- Qin T, Dang Y, Lin T, et al. Single-atom Ru catalyst for selective synthesis of 3-pentanone via ethylene hydroformylation. Green Chem 2021; 23: 9038-9047. [Article] [CrossRef] [Google Scholar]
- Wei X, Jiang Y, Ma Y, et al. Ultralow-loading and high-performing ionic liquid-immobilizing rhodium single-atom catalysts for hydroformylation. Chem Eur J 2022; 28: e202200374. [Article] [Google Scholar]
- Tang P, Paganelli S, Carraro F, et al. Postsynthetic metalated MOFs as atomically dispersed catalysts for hydroformylation reactions. ACS Appl Mater Interfaces 2020; 12: 54798-54805. [Article] [CrossRef] [PubMed] [Google Scholar]
- Ding S, Hülsey MJ, An H, et al. Ionic liquid-stabilized single-atom Rh catalyst against leaching. CCS Chem 2021; 3: 1814-1822. [Article] [CrossRef] [Google Scholar]
- Kresse G, Furthmüller J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys Rev B 1996; 54: 11169-11186. [Article] [CrossRef] [PubMed] [Google Scholar]
- Gajdoš M, Hummer K, Kresse G, et al. Linear optical properties in the projector-augmented wave methodology. Phys Rev B 2006; 73: 045112. [Article] [CrossRef] [Google Scholar]
- Perdew JP, Burke K, Ernzerhof M. Generalized gradient approximation made simple. Phys Rev Lett 1996; 77: 3865-3868. [Article] [CrossRef] [Google Scholar]
- Blöchl PE. Projector augmented-wave method. Phys Rev B 1994; 50: 17953-17979. [Article] [CrossRef] [PubMed] [Google Scholar]
- Mizuno N, Misono M. Heterogeneous catalysis. Chem Rev 1998; 98: 199-218. [Article] [CrossRef] [PubMed] [Google Scholar]
- Talib SH, Yu X, Yu Q, et al. Non-noble metal single-atom catalysts with phosphotungstic acid (PTA) support: A theoretical study of ethylene epoxidation. Sci China Mater 2020; 63: 1003-1014. [Article] [Google Scholar]
- Brown GM, Noe-Spirlet MR, Busing WR, et al. Dodecatungstophosphoric acid hexahydrate, (H5O2+)3(PW12O403−). The true structure of Keggin’s ‘pentahydrate’ from single-crystal X-ray and neutron diffraction data. Acta Crystlogr B Struct Crystlogr Cryst Chem 1977; 33: 1038-1046. [Article] [CrossRef] [Google Scholar]
- Yates Jr. JT, Duncan TM, Worley SD, et al. Infrared spectra of chemisorbed CO on Rh. J Chem Phys 1979; 70: 1219-1224. [Article] [NASA ADS] [CrossRef] [Google Scholar]
- Ding S, Guo Y, Hülsey MJ, et al. Electrostatic stabilization of single-atom catalysts by ionic liquids. Chem 2019; 5: 3207-3219. [Article] [CrossRef] [Google Scholar]
All Tables
All Figures
Figure 1 (A) SEM image of Rh1/Cs-PTA. (B) Tertiary structure: porous particle (100–200 nm). (C) Secondary structure: ionic crystals (10–20 nm). (D) Primary structure: top view of the optimized geometry of Rh1/PTA, ~1 nm), purple: Rh, blue: W, red: O, pink: P. (E) Charge density difference of Rh1/PTA. (F) Spin-polarized partial density of states projected on Rh s, p and d (green, blue and red) states. |
|
In the text |
Figure 2 (Upper) HK-plot analysis of N2 adsorption-desorption isotherms. (Lower) XRD patterns of Rh1/H-PTA (a), Rh1/K-PTA (b), Rh1/Rb-PTA (c), and Rh1/Cs-PTA (d). |
|
In the text |
Figure 3 (A–C) EDX mapping images of Rh1/Cs-PTA. (D) CO DRIFT spectra of different catalysts: Rh/K-PTA, Rh/Rb-PTA, Rh/Cs-PTA, and Rh/Cs-PTA. |
|
In the text |
Scheme 1 Reaction route of the hydroformylation of styrene and other substrates. |
|
In the text |
Current usage metrics show cumulative count of Article Views (full-text article views including HTML views, PDF and ePub downloads, according to the available data) and Abstracts Views on Vision4Press platform.
Data correspond to usage on the plateform after 2015. The current usage metrics is available 48-96 hours after online publication and is updated daily on week days.
Initial download of the metrics may take a while.