Introduction

Supported metal catalysts occupy a pivotal position in diverse technological applications, especially for the sustainable production of chemicals and fuels1,2,3. The increase of metal dispersion on supports by downsizing metals to nanoscale is of significance for maximizing the metal atom utilization and consequently increasing the activity normalized to mass4,5,6,7. The extreme metal utilization can be achieved when the particle size approaches 1 nm (that is, nanoclusters of several dozen atoms), as almost all metal atoms at this size are accessible for catalysis8. On the other hand, the electronic properties of metals are also highly dependent on their size particularly below 2 nm, due to the quantum size effect9, which has been shown to be crucial for enabling the metal catalysts with high intrinsic activity and unexpected selectivity6,10,11,12.

However, an inherent problem of metal nanoclusters for catalysis applications is their well-documented thermodynamic instability13,14,15, as metal species tend strongly to grow into larger crystallites due to the sharply increased surface energy with the decrease of particle size, through the particle migration and coalescence (PMC) and/or Ostwald ripening (OR) processes16,18,19 and the encapsulation of metal nanoclusters with porous nanoshells (e.g., zeolite and metal oxides)20,21,22,23. Albeit conceptually efficient, these methods may cause decreased active surface area and increased mass-transfer resistance, which can considerably lower the overall performance of the catalysts. Recently, some nanostructured metal oxide supports have been proven to be capable of stabilizing metal species at high temperature through metal–support bonding (for example, the Pt–O–Ce bonding in Pt/CeO2 catalyst) in oxidative atmospheres24,25,26,27. In another breakthrough work, Hu et al. reported the high-temperature shockwave method to stabilize metal single atoms on carbon, C3N4, and TiO2 supports by forming the metal-defect bonds28.

Here, we demonstrate that sulfur—a traditional poison reagent for metal catalysts29—when doped in a carbon matrix can efficiently stabilize ~1 nm metal nanoclusters (Pt, Ru, Rh, Os, and Ir) against thermal sintering at high temperatures up to 700 °C in a reductive atmosphere. Spectroscopic characterization and density functional theory (DFT) calculations reveal that the enhanced adhesion strength at the metal/sulfur-doped carbon (S–C) interface, arising from the strong and thermally stable metal-sulfur bonding, greatly suppresses the metal nanocluster sintering kinetics by retarding both metal atom diffusion (OR path) and nanoparticle migration (PMC path). Moreover, the prepared Pt nanocluster catalysts exhibit high activity, selectivity, and long reaction lifetimes in the high-temperature reaction of propane dehydrogenation (PDH) to propylene at 550 °C, demonstrating the application potentials of the metal nanocluster catalysts for industrially relevant catalysis under realistic technical conditions.

Results

Sinter resistance of S–C supported nanocluster catalysts

Figure 1 shows our concept of suppressing metal migration/coalescence and ripening to achieve the sinter-resistant metal nanocluster catalysts based on the strong chemical interaction between metal and doped sulfur atoms in the S–C supports. We first prepared the S–C support by cobalt-assisted carbonization of molecular precursors (that is, 2,2’-bithiophene) with SiO2 nanoparticles as templates, according to our previously reported works30,31. The S–C support with structural features of a high sulfur content (~14 wt %) and a large specific surface area (>1200 m2g−1) then served as the support for loading of Pt nanoclusters (1 wt %) by the wet-impregnation of H2PtCl6 and subsequent H2-reduction at 700 °C. High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) measurements of the S–C supported Pt nanocluster (Pt/S–C) catalyst showed that numerous Pt nanoclusters with size in the range of 0.8–1.3 nm (mean size, 1.0 nm) homogeneously distributed over the whole carbon matrix, without any large particles or nanocluster aggregates (Fig. 2a and Supplementary Fig. 1). Aberration-corrected HAADF-STEM image with atomic resolution indicates that the Pt nanoclusters are highly crystalline and mainly enclosed by (111) and (200) crystal planes of face-centered cubic Pt (Supplementary Fig. 2). Note that there were still considerable sulfur atoms survived in the carbon carrier even after the high-temperature H2-reduction at 700 °C during catalyst preparation (Fig. 3a). We further surmise that these thermally stable sulfur atoms are crucial for stabilizing metal nanoclusters at high temperatures.

Fig. 1: Schematic illustration of sulfur-stabilizing method.
figure 1

The S–C supported metal nanocluster catalysts are resistant to sintering owing to the strong metal-sulfur interaction, whereas S-free-C supported catalysts easily aggregate into larger particles in high-temperature reactions.

Fig. 2: Sintering tests.
figure 2

ah HADDF-STEM images of the catalysts before (a, c, e, g) and after (b, d, f, h) sintering tests, including 1 wt% Pt/S–C (a, b), 5 wt% Pt/S–C (c, d), 1% Pt/S-free-C (e, f), and commercial Pt/C catalyst (g, h). Sintering test condition: 5% H2/Ar, 700 °C, 600 min.

Fig. 3: Spectroscopic characterizations.
figure 3

a XPS survey spectra of the S–C support and the Pt/S–C catalyst before and after annealing treatment at 700 °C in 5% H2/Ar for 600 min. b Pt L3-edge XANES spectra of Pt/S–C, Pt/S-free-C, PtS2, and Pt foil. c Fourier transform of the EXAFS spectra of Pt/S–C, Pt/S-free-C, PtS2, PtO2, and Pt foil. d In-situ XPS of Pt 4f on Pt/S–C and Pt/S-free-C after H2 treatment at 200 °C for 30 min.

Owing to the dramatically increased surface energy of Pt particle below 2 nm compared to their bulk counterparts, the small-sized Pt nanoparticles would be highly mobile and have strong tendency of migration/coalescence and ripening even on high-surface-area supports8). In the case of commercial Pt/C catalyst, as expected, we also observed the severe aggregation and overgrowth of Pt nanoparticles after the thermal treatment (Fig. 2g and h), due to the lack of strong metal-carbon support interactions.

To demonstrate the wide applicable scope of our synthetic strategy, we additionally prepared another four platinum-group metal (PGM) nanocluster catalysts and studied their anti-sintering properties under the same conditions, including Ru, Rh, Os, and Ir (1.0 wt% loading for all cases). HAADF-STEM images showed that these catalysts were composed of highly dispersed metal nanoclusters with mean sizes in the range of 0.46–0.82 nm (Supplementary Fig. 9). Low-magnification HAADF-STEM images with large views further indicated the homogeneity of these metal nanoclusters (Supplementary Fig. 10). Importantly, these metal nanoclusters also exhibited remarkable sinter-resistant properties without visible particle growth and agglomeration during the long-term thermal treatment at 700 °C for 10 h (Supplementary Figs. 9, 10) and the sulfur still existed stably in the carbon matrix (Supplementary Fig. 11 and Supplementary Table 1). The S-free-C supported Ru, Rh, Os, or Ir nanoclusters, in marked contrast to S–C supported ones, experienced different degrees of sintering; the average size of Ru, Rh, Os, and Ir nanoclusters on the S-free-C supports increased greatly from below 1 nm to 13.74, 7.96, 5.24, and 6.12 nm, respectively (Supplementary Figs. 1214). Moreover, the sulfur-stabilization method was further extended to prepare stable bimetallic and multi-metallic cluster catalysts (Supplementary Fig. 15) and even to prepare sintering-resistant nanocluster catalysts with sulfur-doped TiO2 support (Supplementary Figs. 16, 17).

Interfacial electronic effects between Pt and S–C

We then performed X-ray absorption near edge structure (XANES) and extended X-ray absorption fine structure (EXAFS) analyses to verify the strong metal/sulfur interactions. The Pt/S–C catalyst exhibited a considerably lower intensity of white line peak in the Pt L3-edge XANES spectra than Pt/S-free-C (Fig. 3b), even though both of them had a similar particle size (Fig. 2a, e). The white line intensity directly reflects the unoccupied density of states of Pt 5d orbitals37. The lower intensity of white line peak for Pt/S–C suggests that the S–C support can donate electrons to Pt nanoclusters, which leads to an electron-enriched state of Pt nanoclusters on S–C. Such electronic interaction between Pt and S–C support has been confirmed previously by the density functional theory (DFT) calculations38.

In the Fourier transforms of EXAFS data, Pt/S–C showed two prominent peaks at ~1.88 Å from the contributions of Pt binding with non-metallic element and ~2.5 Å from the Pt–Pt contributions, respectively (Fig. 3c). By comparing with the EXAFS of PtS2 and PtO2, we can safely conclude that the peak of 1.88 Å observed in Pt/S–C is exclusively associated to the Pt–S bonds, again implying the strong chemical interaction between Pt and S–C (Fig. 3c and Supplementary Fig. 18). Note that the Pt–Pt bonding distance in Pt/S–C is shortened than bulk platinum (~2.7 Å) owing to the existence of small Pt nanoclusters with low-lying electronic states in the range of 2.4–2.6 Å as revealed by theoretical studies

Fig. 4: Theoretical investigations of nanoparticle diffusion and atom esca**.
figure 4

a Energy barrier of Pt38 cluster desorption on S–Graphene and Graphene. b Escape energy of one selected atom (site 4) from the Pt38 cluster on S–Graphene and Graphene. c The selection of several other atoms at different positions for the escape energy calculations. d The escape energy of each atoms at different positions shown in (c). The higher energy barriers for both Pt38 migration and individual atom esca** on the S–Graphene substrate support the experimentally observed sinter resistance of Pt nanoclusters on S–C.

To understand the direction of charge transfer between metal clusters and the S–C support, we calculated the valence band maximum (VBM) level of S–Graphene and the conduction band minimum (CBM) of Pt38 cluster by using the Frontier Orbital Theory. Due to the VBM level of S–Graphene is higher than the CBM level of Pt38 cluster, electrons will transfer from S–Graphene to Pt38 cluster under the interaction of two orbitals, where the Pt38 cluster spontaneously donate empty orbitals to capture electrons (Supplementary Fig. 19a). Moreover, we further explored the Bader charge analysis and found that Pt38 cluster could capture 0.73 electrons from the S–Graphene (Supplementary Fig. 19b), which further demonstrated an electron-enriched state of Pt nanoclusters on S–C by the interfacial metal–sulfur electronic interactions.

For the simulation of sintering process, we considered both PMC and OR sintering paths. The PMC sintering mechanism involves the mobility of the whole particles on the support surface with Brownian motion collide16. In this case, the diffusion of particle needs to overcome the strength of adhesion at the metal/support interface. As shown in Fig. 4a, our DFT calculations suggested a much higher energy barrier of the Pt38 nanocluster desorption on S–Graphene (3.49 eV) than that on Graphene (0.63 eV), indicating that the PMC process of Pt nanoclusters can be significantly suppressed on the S–C supports. The OR process involves the migration of individual atom driven by the differences of surface energy of nanoparticles, in which the atomic species are emitted from one nanoparticle, diffuse over the supports and attach to another nanoparticle, resulting in the gradual growth of larger particles and consumption of small particles16. By DFT simulation, we found that the escape energy of individual atom from Pt38 cluster on S–Graphene was larger than that on Graphene (Fig. 4b–d), which also can be ascribed to the enhanced strength of interfacial adhesion at the Pt38/S–Graphene owing to the Pt–S bonding.40 The increased adhesion energy would greatly reduce the chemical potential of the surface metal atoms17, which eventually suppress the inter-particle sintering kinetics by retarding the Ostwald ripening pathway. Overall, these calculation results clearly demonstrate that the strong metal/S–C interaction can dramatically enhance the anti-sintering ability by inhibiting both metal atom diffusion (OR path) and nanoparticle migration (PMC path).

Catalytic propane dehydrogenation

Taking into account the outstanding thermal stability of the S–C supported metal nanoclusters, we tested the catalysis application of Pt/S–C for the industrially important propylene production by PDH. It is well-documented that PDH is highly endothermic and thermodynamically limited and therefore requires a high reaction temperature, typically in the range of 400–600 °C41. Under such harsh conditions, it is extremely challenging to maintain a high propene selectivity and suppress the irreversible deactivation caused by undesirable metal nanoparticle sintering and heavy coking11.

The PDH on Pt/S–C was first studied in a fixed-bed micro-reactor at 550 °C with 10 vol% C3H8 feeding. Pt/S-free-C, commercial Pt/C, Pt/Al2O3, and PtSn/Al2O3 catalysts were also tested under the same conditions. Notably, the Pt/S–C catalyst exhibited 40% conversion corresponding to a high initial propylene formation rate of 2028 mmol gPt−1 h−1 (Fig. 5a), which only slightly decreased to 1986 mmol gPt−1 h−1 (corresponding to 2.1% deactivation) after continuous operation for 600 min (Fig. 5b). The long-term durability of Pt/S–C for 1800 min operation further demonstrated its outstanding stability (Fig. 5c). In contrast, the propylene formation rate on the Pt/S-free-C catalyst decreased greatly decreased from 1745 to 1071 mmolgPt−1 h−1 (corresponding to 38.6% deactivation) after 600 min (Fig. 5a, b). Although Pt/Al2O3 exhibited a higher initial propylene formation rate (3303 mmol gPt−1 h−1) than Pt/S–C, the rate rapidly decayed to 1658 mmol gPt−1 h−1 after 120 min, which further decayed to 1340 mmol gPt−1 h−1 after 600 min (corresponding to 59.4% deactivation). The commercial Pt/C catalyst exhibited a much lower activity with initial rate of only 570 mmol gPt−1 h−1 as well as a poor stability with 73.5% deactivation (Fig. 5a, b). Besides, the conversion on PtSn/Al2O3 catalyst gradually decreased from 39 to 32% (corresponding to 17.9% deactivation) after 600 min (Fig. 5a), then further slowly decayed to 30% after 950 min (Supplementary Fig. 21). The first-order deactivation model was used to further analyze the catalysts stability. The much lower deactivation rate of 0.005 h−1 for Pt/S–C quantitatively demonstrates its superior stability compared to Pt/S-free-C (0.05 h−1), Pt/C (0.15 h−1), Pt/Al2O3 (0.166 h−1), and PtSn/Al2O3 (0.031 h−1). Meanwhile, when the C3H8 concentration in feeding was increased from 10 vol% to 33 vol% and 50 vol%, although the conversion on Pt/S–C decreased from 40% to 19% and 13%, respectively, the Pt/S–C catalyst still showed a high stability with a comparatively lower deactivation rate of 0.003 h−1 at 33 vol% C3H8 feeding and 0.005 h−1 at 50 vol% C3H8 feeding. In contrast, for the Pt/S-free-C reference catalyst, the C3H8 conversion dramatically dropped from 30 to 18% with the increase of propane concentration (Supplementary Fig. 22). In addition, with the increases of weight hourly space velocity (WHSV) and temperature (Supplementary Fig. 23), the Pt/S–C catalyst could achieve a very high propane rate of 6645 mmol gPt−1 h−1.

Fig. 5: Catalytic performance for the high-temperature PDH.
figure 5

a C3H8 conversion of Pt/S–C, Pt/S-free-C, commercial Pt/C, Pt/Al2O3 and PtSn/Al2O3 catalysts as a function of time during PDH. b Propylene formation rate at the initial stage and after 600 min continuous PDH operation. c Long-term stability test of Pt/S–C for PDH at 550 °C for 1800 min. d C3H6 selectivity of Pt/S–C, Pt/S-free-C, commercial Pt/C, Pt/Al2O3 and PtSn/Al2O3 catalysts as a function of time during PDH.

Moreover, the Pt/S–C catalyst showed a high selectivity of ~98% in the whole catalysis process under 10% C3H8 concentration feeding (Fig. 5d), which is comparable to PtSn/Al2O3 and higher than the selectivity of Pt/S-free-C (~90%), Pt/C (90%), and Pt/Al2O3 (17–70%). The low selectivity of Pt/Al2O3 relative to carbon supported Pt catalysts is associated to the strong acidic sites of Al2O3, which promotes side reactions and coke formation42. Further, the Pt/Al2O3 could rapidly become deactivated through the blocking of the metal active site by coke deposition43. HAADF-STEM measurements was employed to monitor the morphology change of these catalysts before and after the PDH reaction. As speculated, serious sintering of the Pt nanoclusters happened on Pt/S-free-C, Pt/C, Pt/Al2O3, and PtSn/Al2O3, while the size and morphology of the Pt nanoclusters in Pt/S–C were essentially unchanged (Supplementary Figs. 24, 25). Under high C3H8 concentration feedings of 33% and 50%, the selectivity of Pt/S-free-C showed a sharp decline from 90% to 80% and 50% and, therefore, exhibited rapid deactivation, while the selectivity of Pt/S–C showed slight decrease from 98% to 94% and 88% but still with a stable activity (Supplementary Fig. 26). When the reaction temperature rose to 600 °C, the Pt/S–C catalyst exhibited a slightly worse selectivity of around 92% than that under 550 °C with the decreased conversion from 55% to 40% after 1200 min, whereas the Pt/S-free-C catalyst showed a rapid deactivation at the beginning of the reaction (Supplementary Fig. 27). The widened performance gap between these two catalysts at more harsh conditions further demonstrated the pivotal role of the Pt–S interactions in promoting the selectivity and durability in catalyzing PDH.

We further compared the performance of the Pt/S–C catalyst with the reported monometallic Pt, Pt alloy, as well as metal oxide catalysts, in terms of the space time yield (STY) of propene formation, the selectivity to propene, and deactivation rate (Supplementary Fig. 28 and Table S3). At high C3H8 feed, the Pt/S–C catalyst exhibited a lower C3H6 STY than some metal oxides catalysts and most Pt alloy catalysts and acquired a higher C3H6 STY under high WHSV (5 h−1) at 10 vol% C3H8 feed, which is superior to most monometallic Pt and even some Pt alloy catalysts. As for the selectivity, there is little difference between Pt alloy catalysts and the Pt/S–C (above ~90% for both kinds of catalysts), while some metal oxide and monometallic Pt catalysts showed low selectivity of 75–85%. In the term of deactivation rate, the Pt/S–C is obviously in the top position with outstanding stability and even better than most Pt-alloys catalysts.

Considering the relatively low activity of Pt/S–C under 50 vol% C3H8 feed, we further synthesized bimetallic Pt–Sn/S–C catalysts with the target of further promoting C3H6 STY and selectivity meanwhile maintaining low deactivation rate at high C3H8 feed (Supplementary Fig. 29). The bimetallic Pt–Sn/S–C catalyst exhibited an enhanced activity with a conversion of 25% and also a promoted selectivity of 94% compared to monometallic Pt/S–C catalysts (13% conversion and 88% selectivity) under the identical reaction conditions. Meanwhile, the stability was well maintained with a low deactivation rate constant of 0.003 h−1 for 3000 min. These results make the Pt–Sn/S–C catalyst be one of the best PDH catalysts (Supplementary Fig. 28 and Table S2). The systematic optimization of the Pt–Sn/S–C catalyst and comprehensive screen of bimetallic combination of alloys may further improve the performance, but it has seriously deviated from the core concept of this work.

We also tried to apply the sulfur-doped TiO2 support to regeneration for PDH reaction under oxidation conditions. After 5 successive regeneration cycles, the performance remained little change (Supplementary Fig. 30), indicating the feasibility of anti-sintering under oxidation conditions of metal–S interaction on industrially relevant metal oxide supports. The remarkable sintering resistance of the Pt/S–C catalyst highlights its potential technological value in realistic catalysis environments. For example, for the water-gas shift (WGS) reaction, Pt/S–C also exhibited a stable catalytic performance after continuous operation for 3500 min (Supplementary Fig. 31).

We carried out in-situ thermogravimetric analysis (TGA) under reactive atmosphere of PDH to directly monitor the coke formation at 550 °C (Fig. 6a). Encouragingly, the Pt/S–C exhibited outstanding anti-coking ability with no any detectable coke accumulation under the long-time reaction, indicating the easier propylene desorption on electron-enriched Pt nanoclusters on the S–C support. In sharp contrast, coke formation on the Pt/S-free-C catalyst was heavy, reaching a 6.2% of weight gain after 600 min, which is ascribed to the strong C3H6 adsorption and deep dehydrogenation on Pt/S-free-C. The above XANES, EXAFS, and XPS have demonstrated that the Pt nanoclusters were in an electron-enriched environment induced by the electron donation from the S–C support to Pt. The increased electron density of Pt nanoclusters can weaken the bond between Pt and electron-rich group of C=C of propylene on account of electrostatic repulsion44, which would enhance the propylene desorption and accordingly improve the PDH selectivity. An additional advantage of the enhanced propylene desorption is to efficiently avoid the further dehydrogenation of propylene and eventually the coke formation (Fig. 6b), which is another deactivation mechanism in addition to the metal particle sintering45.

Fig. 6: Experimental and theoretical interpretation of propylene selectivity.
figure 6

a In-situ TGA of Pt/S–C and Pt/S-free-C under reactive atmosphere of PDH. b Schematic illustration of the crucial role of interfacial electronic effects between Pt and S–C in promoting the selectivity and stability of PDH reaction. c Calculated energy barriers for dehydrogenation steps of propane on Pt38/S–Graphene, Pt38/Graphene, and Pt (111). The dotted lines indicate the desorption barriers of C3H6* to gaseous C3H6.