Abstract
Catalytic activity of metal particles is reported to originate from the appearance of nonmetallic states, but conductive metallic particles, as an electron reservoir, should render electron delivery between reactants more favorably so as to have higher activity. We present that metallic rhodium particle catalysts are highly active in the low-temperature oxidation of carbon monoxide, whereas nonmetallic rhodium clusters or monoatoms on alumina remain catalytically inert. Experimental and theoretical results evidence the presence of electronic communications in between vertex atom active sites of individual metallic particles in the reaction. The electronic communications dramatically lower apparent activation energies via coupling two electrochemical-like half-reactions occurring on different active sites, which enable the metallic particles to show turnover frequencies at least four orders of magnitude higher than the nonmetallic clusters or monoatoms. Similar results are found for other metallic particle catalysts, implying the importance of electronic communications between active sites in heterogeneous catalysis.
Similar content being viewed by others
Introduction
A majority of catalytic oxidation reactions often involve the catalyst-mediated electron transfer between reactants1,2,3,4, in which a catalyst acts as an electron reservoir by accepting electrons from one reactant and then delivering them to other reactant(s)5, and thus the electronic states of the catalyst, particularly in terms of nonmetallic states or metallic states, have an important influence on reaction rates6,7,8,9,10. There exist two prevailing yet diametrically opposed views on the function of the electronic states, emphasizing the importance of either nonmetallic states6,7 or metallic states8,9,10 in determining catalytic performance. The fundamental discrepancy between these views lies in whether there exists electronic communications between active sites when a reaction takes place11. Nonmetallic states lead to a lack of electronic communications, so catalytic reactions merely occur locally on individual active sites. In contrast, metallic states can give access to electronic communications for active sites, thus opening up an electron-delocalized mechanism (EDM)11.
In order to test both views, an elaborately designed system including a prototypical reaction and model catalysts is necessary. Carbon monoxide (CO) oxidation is commonly used as a typical probing reaction, yet it is significantly important in industrial catalysis12. According to the Sabatier principle on medium adsorption strength13, metals with not fully unfilled frontier dn (particularly, 5 < n < 10) orbitals are often favorable for effectively adsorbing and activating CO and O2 molecules14,15, and thus rhodium (Rh) with an electronic configuration of d8s1 at its ground state should be a suitable candidate16. As the metal atomicity increases, there often exists a nonmetallic-to-metallic state transition region6,17,18,19,20, and, as typical representatives for metallic and nonmetallic catalysts, two reliable Rh catalyst models should require their atomicity respectively to approach two sides of this transition region. A catalytically inert and electrically insulating support such as Al2O3 is also desired to disperse and stabilize Rh clusters or nanoparticles21, and to eliminate or rule out the so-called electronic metal-support interaction22,23.
In this work, we provide direct evidence that electronic communications are present between active sites on individual metallic particles, which enable metallic rhodium particles to possess four orders of magnitude higher turnover frequency than that of nonmetallic particles without such communications. Experimental and theoretical results demonstrate that inter-site electron communications on the metallic particles drive two half-reactions coupling between active sites, thus allowing CO oxidation to proceed along a lower-activation-energy kinetic pathway than the nonmetallic clusters or monoatoms.
Results
Design of metallic and nonmetallic model catalysts
To achieve metallic Rh nanoparticles with desired sizes and well-defined nanofacets24, a colloid preparation method is used to obtain colloidal Rh nanoparticles. These Rh nanoparticles are then deposited onto Al2O3 surfaces to gain a metallic Rhmet/Al2O3 catalyst. For comparison, a nonmetallic Rhnon/Al2O3 catalyst is prepared by a conventional impregnation route (See Supplementary Figs. 1 and 2, Supplementary Table 1, and Methods for details). The metallic and nonmetallic states of Rhmet and Rhnon are verified by the valance band X-ray photoelectron spectra (VB XPS)25. In Fig. 1a, Rhmet/Al2O3 shows a significant electronic state density on the Fermi level (EF), evidencing that the Rh nanoparticles are in metallic states, whereas the electronic state density of Al2O3 appears at approximately 3.2 eV below EF (Supplementary Fig. 3)26. For comparison, the electronic state density nearest to EF for Rhnon/Al2O3 is at −0.6 eV (Fig. 1b), corresponding to nonmetallic states. These results are consistent with the X-ray absorption near-edge structure data (Supplementary Fig. 4).
a, b VB XPS of Rhmet/Al2O3 (a) and Rhnon/Al2O3 (b). A red shade in (a) shows the electronic state density near EF (the Fermi level) for Rhmet/Al2O3, and onset of the electronic state density of Rhnon/Al2O3 is analyzed to appear at −0.6 eV in (b). The curves were slightly smoothed. c, d AC-STEM images of Rhmet/Al2O3 (c) and Rhnon/Al2O3 (d). Rh nanoparticles in (c) and Rh clusters in (d) are selectively circled with yellow dashed lines. Insets: the size distribution of Rh particles or clusters. e, f Atom-resolution AC-STEM images of Rhmet/Al2O3 (e) and Rhnon/Al2O3 (f). Insets: two Wulff models in (e) are constructed according to the corresponding image of one typical Rh nanoparticle closed by white dashed lines (The gold, purple, and green balls represent Rh atoms at vertex sites, edge sites, and on the nanofacets) and 12-atom raft models on the Al2O3 surface are also drawn on the basis of the image of one Rh cluster in (f). g, h WT EXAFS spectra of Rhmet/Al2O3 (g) and Rhnon/Al2O3 (h) at Rh K-edge.
The electronic states of metal particles are intimately associated with particulate sizes on the (sub)nanometer scale. Fig. 1c, d displays the aberration-corrected scanning transmission electron microscopy (AC-STEM) images of Rhmet/Al2O3 and Rhnon/Al2O3, respectively. The Rh particles of Rhmet/Al2O3 exhibit sizes of 2–3 nm, with an average size of 2.3 nm (Supplementary Fig. 5 and inset in Fig. 1c). A typical Rh nanoparticle together with a corresponding Wulff model (~420 atoms) is shown in Fig. 1e. The Rh particles in this size range are evidenced to possess the metallic behavior (Fig. 1a), consistent with the experimental finding that metal particles with sizes larger than 1.5 nm are in metallic states6,20. On the other hand, the average size of the Rh clusters in Rhnon/Al2O3 is ~0.9 nm (Fig. 1d, f, and Supplementary Fig. 6). Inasmuch as the nonmetallic-to-metallic state transition for some metal particles, including Rh particles, often occurs in a size range of 1–2 nm6,18,19,20, Rhmet/Al2O3 and Rhnon/Al2O3 with the Rh sizes on two sides of the state transition region can act as two reliable representatives for metallic and nonmetallic catalysts, respectively.
Owing to the low work function of Rh (5.5 eV)27,28, O2 dissociation adsorption energies (Ead) on Rh particles are large (Supplementary Fig. 7). Therefore, Rh particles are often covered by one surface oxide layer15. O2 dissociation adsorption on Rh particles is evidenced for both Rhmet/Al2O3 and Rhnon/Al2O3, as shown by the Rh-O bonds in Fourier-transformed extended X-ray absorption fine structure (FT EXAFS) spectra (Supplementary Fig. 8) and wavelet transform (WT) EXAFS spectra in Fig. 1g, h, thus leading to the partial positively charged Rh atoms via donating electrons to the adsorbed O species (Supplementary Fig. 9). Likewise, upon CO adsorption, surface Rh atoms will accept electrons from adsorbed CO molecules29. As for one metallic particle, the electron extraction or injection occurring on one active site enables the charges to be redistributed on the entire metallic particle, leading to electronic communications with another entangled active site, whereas such site-site electronic communications do not exist on nonmetallic particles in principle. Hence, Rhmet/Al2O3 and Rhnon/Al2O3 are reliable model catalysts to test both views aforementioned.
Evidence for inter-site electronic communications
With two model catalysts, we investigate the possibility of electronic communications between active sites on individual metallic particles in CO oxidation. Although Rhnon/Al2O3 has much more catalytic sites than Rhmet/Al2O3 in virtue of smaller particles exposing more Rh atoms on surfaces, Rhmet/Al2O3 shows much higher activity in CO oxidation than Rhnon/Al2O3 under identical conditions (Fig. 2a and Supplementary Fig. 10). CO oxidation on Rhmet/Al2O3 starts at 60 oC and reaches 100% conversion at 130 oC, at which Rhnon/Al2O3 remains catalytically inactive. Fig. 2b displays apparent activation energies (Ea) extracted from the Arrhenius plots (Supplementary Fig. 11), and the Ea value of Rhmet/Al2O3 is ~40 kJ mol−1, in line with the reported value for CO oxidation over Rh nanoparticle catalysts30. A large difference (60 kJ mol−1, or ~0.6 eV) in Ea between Rhmet/Al2O3 and Rhnon/Al2O3 coincidentally equals minimal energy required for an electron transition from the ground state to EF of Rhnon/Al2O3 (Fig. 1b), implying the validity of the interfacial energy-alignment principle31,32 that a molecule’s HOMO (highest occupied molecular orbital) level can become pinned to the catalyst’s EF level on which to exchange the charge. Considering the discrepancy in electronic state between two catalysts, we infer that the high activity and the low Ea of Rhmet/Al2O3 are closely related to the metallic states.
a CO conversions as a function of temperature over Rhmet/Al2O3 and Rhnon/Al2O3. b Average Ea values with the error bars of CO oxidation over Rhmet/Al2O3 and Rhnon/Al2O3. c O2-TPD profiles of Rhmet/Al2O3 and Rhnon/Al2O3 after O2 saturation adsorption at 50 oC. DO2 represents the desorption amount of O2. The red shade stands for the low-temperature desorption peak of O2 from Rhmet/Al2O3. d Intensity contours of in situ DRIFT spectra of Rhnon/Al2O3. The bands at ~2090 and 2018 cm-1 are assigned to symmetric (vs) and asymmetric (vas) stretching vibration modes of the adsorbed CO molecules on the edge Rh atoms, respectively. The band appearing at ~2116 cm−1 is due to O2−-Rh2+-CO species. Inset: adsorption model on individual Rh atom, and the purple, black, and red balls represent Rh, C, and O atoms, respectively. e CO-TPD profiles of Rhmet/Al2O3 after CO saturation adsorption at 50 oC. DCO represents the desorption amount of CO. Inset: a Wulff model of one Rh nanoparticle. The gold, purple, and green balls represent Rh atoms at vertex sites, edge sites, and on the nanofacets, respectively. f, Intensity contours of in situ DRIFT spectra of Rhmet/Al2O3. g Excess electronic charge of CO (δQCO) and the oxidation state (n+) of Rh (Rhn+) as a function of square adsorbed CO vibration frequency (ν2). These data are taken from the refs. 33,37. h CO DRIFT spectra on Rhmet/Al2O3. The spectrum is collected at 30 oC after CO saturation adsorption in a flow of He, and then the sample is heated to 150 oC in a flow of O2/He to collect the spectrum.
One important feature of the metallic states is that the valence band and the conduction band coincide on EF, where to extract or inject electrons thus becomes easy14. The formation of the surface oxide layer (due to O2 + 4e− → 2O2−) reflects that electrons extracted from EF of the metallic Rh particles are energetically favorable (Fig. 1g, and Supplementary Fig. 8). In the O2 temperature programmed desorption (O2-TPD) procedure profile of Fig. 2c, a low-temperature O2 desorption peak at 145 oC due to O2− → ½O2 + 2e− indicates the electron injection onto EF of the metallic Rh particles, whereas such an electron injection is rather difficult for the nonmetallic Rh clusters judging from the neglected O2 desorption amount from Rhnon/Al2O3. Likewise, a new band ( ~ 2116 cm−1) appears at temperatures higher than 250 oC in in situ diffuse reflectance infrared Fourier transform (DRIFT) spectra of Rhnon/Al2O3 (Fig. 2d and Supplementary Fig. 12), which is assigned to the C = O vibration of O2−-Rh2+-CO species (one important reaction intermediates)33. The existence of O2−-Rh2+-CO indicates that the reaction between co-adsorbed CO and O2− on individual electrically insulating Rh atoms (CO + O2− → CO2 + 2e−) would not occur, implying that it is difficult to inject electrons to the nonmetallic Rh clusters.
The identification of active sites is an essential prerequisite for studying site-site electronic communications. The oxide layers on the metallic particle nanofacets are evidenced to be a hexagonal O-Rh-O trilayer15, which are catalytically inactive in low-temperature CO oxidation34, whereas the adsorbed O species from the edge Rh atoms are active and readily desorbed to form under-coordinated sites (Fig. 2c). Subtly, these under-coordinated sites consist of the vertex Rh atoms and the edge Rh atoms not including ending atoms with an atomic ratio of ~1:8 (inset of Fig. 2e, and Supplementary note 1), which can be distinguished by using CO temperature-programmed desorption (CO-TPD) procedure. In Fig. 2e, two CO desorption peaks (denoted as PL and PH) are observed with a calculated PL/PH ratio in desorption amount of ~1:8, which virtually equals the ratio of the vertex Rh atoms to the edge atoms, and thus PL and PH can be attributed to CO desorption from the vertex Rh atoms and the edge atoms, respectively. Obviously, the adsorbed CO molecules on the edge Rh atoms are more stable, and remain catalytically inert in the CO oxidation atmosphere even at temperatures higher than 130 oC (Fig. 2f) (see also discussion in Supplementary note 2), at which CO conversion on Rhmet/Al2O3 reaches 100% (Fig. 2a). Only the CO molecules adsorbed on the vertex Rh atoms are active at low temperatures, evidencing the vertex Rh atoms as active sites.
Every vertex Rh atom on the well-defined-facet Rh particles has identical geometrical and electronic structures, which allows one to acquire site-site electronic communication information on the basis of macroscopic data measured from these uniform active sites by using site-specific techniques35. The electronic communications originate essentially from the charge redistribution of the charged active sites on the entire metallic particle, and, as for a metallic particle with well-defined structure, such a redistributed charge to other vertex active sites has proven to be more than that to the edge atoms36. To verify whether such electronic communications exist on individual Rh metallic particles, we can monitor the electronic perturbations of the edge atoms by the vertex charged active sites. The DRIFT spectroscopy of CO on noble metal catalysts is site-specific and sensitive to electronic perturbations of the binding sites33,37 (Fig. 2g). In Fig. 2h, we first collect a DRIFT spectrum of CO-saturated Rhmet/Al2O3 at 30 oC, and then collect another DRIFT spectrum after heating the sample up to 150 oC to make the vertex active sites positively charged after O2 introduction via substituting the adsorbed CO by surface lattice oxygen (CO + O2 + 2e− → CO2 + O2−). An observed blue-shift by 11 cm−1 (Fig. 2h) elucidates a charge decrease of the edge atoms (Fig. 2g and Supplementary note 3), reflecting the charge redistribution on the metallic particle. These results evidence that there exist electronic communications between active sites on individual metallic particles. The electronic communications can speed up CO oxidation so as to achieve a high turnover frequency (TOF = 11.2 s−1) at 160 oC, at least four orders of magnitude higher than TOF (6.0 × 10−4 s−1) of the nonmetallic clusters or the monoatoms without such electronic communications (Supplementary Fig. 13 and Supplementary note 4).
Mechanisms of electronic-communications-driven reactions
To shed light on reasons why the electronic communications can enhance the reaction rate, we carry out DFT simulations to explore CO oxidation mechanisms over Rhmet/Al2O3. A reliable metallic particle Wulff structure containing thirty-eight Rh atoms (see Supplementary Fig. 7 for details) with nanofacets covered by twenty-four O atoms is established, which exposes twenty-four vertex atoms as active sites on surfaces. Noticing that the calculated adsorption free energies of O and CO on the vertex active sites are almost identical, twelve CO molecules and twelve O atoms, as electron donor and acceptor, respectively, are alternately adsorbed on twenty-four vertex active sites (Supplementary Fig. 14), which allows every active site to have an identical chemical potential such that there is no driving force for electron transfer between active sites, and thus we can randomly choose one of the active sites on which to simulate a catalytic cycle with the local reaction mechanism (LRM)11. To simulate the EDM with inter-site electronic communications11, eleven sites of the twelve adsorbed CO were replaced by adsorbed O atoms (Supplementary Fig. 14), on each of which the half-reaction of ½O2 + 2e− → O2− is assumed to have taken place via extracting two electrons from each active site, thus allowing the metallic particle to be positively charged as the result of the charge redistribution. Therefore, we simulate a catalytic cycle at the reserved CO adsorption site.
The DFT calculation results on LRM and EDM of CO oxidation over Rhmet/Al2O3 are shown in Fig. 3a, in which two kinds of the reaction mechanisms follow a same pathway. There are two critical transition states (TSs) corresponding to the O-CO bond formation steps in the CO oxidation catalytic cycle for both cases (Inset of Fig. 3a and Supplementary Fig. 15), which are involved in the electron-injection and electron-extraction processes: (i) CO + O2− → CO2 + 2e−, and (ii) CO + O2 + 2e− → CO2 + O2−. Owing to the strong intrinsic activation ability of the metallic Rh particles toward O215, the energy required for extracting electrons from the active sites is lower than that for injecting electrons, and the reaction (i) with TS1 is thus the rate-determining step for both cases. In Fig. 3b, the energy barrier (0.57 eV) of the rate-determining step for EDM is lower than that (0.80 eV) for LRM despite sharing the same pathway, and the former agrees satisfactorily with the experimental Ea value (0.42 eV). Thermodynamically, an exothermic process with an energy saving of 0.17 eV for the rate-determining step allows CO oxidation to follow EDM rather than LRM with the endothermic rate-determining step requiring an energy input of 0.27 eV. These results are consistent with the fact that the EDM is commonly present in metallic particle catalysis4,11,38,39. By coupling two redox half-reactions, the inter-site electronic communications can effectively reduce the energy barrier of the rate-determining step, thus lowering Ea and enhancing the overall reaction rate.
a Structural snapshots illustrating the DFT-calculated reaction pathway of CO oxidation on Rhmet/Al2O3 with electron-injection and electron-extraction processes involving six steps for a cycle. The gold and light gold balls represent active site Rh atoms and the other Rh atoms, respectively. The red and light red balls represent O atoms. The black balls represent C atoms. Inset: the relative free-energy diagram for Rhmet/Al2O3 following EDM or LRM. b Energy barriers of the EDM and LRM rate-determining steps of CO oxidation on Rhmet/Al2O3. ΔG represents the Gibbs free energy difference with respect to the initial state (A). c Projected density of state (pDOS) of gaseous CO molecule (left), and adsorbed CO, reactive surface lattice oxygen (O2−) and the vertex Rh active site at the adsorption states (or at A in panel (a) (middle) together with an enlargement of the rectangle closed by the red dashed lines in the middle diagram (right). The red dashed lines serve only as a visual guide in order to show the relationship between the enlargement and the original diagram. The energy level was normalized to the vacuum level and EF represents the Fermi level of Rhmet/Al2O3. The electron number on each orbital is also given. Black dashed lines show CO orbital shifts and redistribution after the CO adsorption. d pDOS of reactive CO on Rhmet/Al2O3 at the adsorption states (or at A in panel (a) and TS1.
We focus on the charge transfer between the active site and the reactants at the rate-determining step (CO + O2− → CO2 + 2e−) for the low-Ea EDM. To produce CO2 with two C = O bonds14,40 from CO with a C ≡ O bond (Fig. 3c), the cleavage of one 1π bond and the depletion of two 5σ electrons of CO are required for forming a new C = O bond. At the adsorption state (Fig. 3c), the d → 2π* back-donation41 by ~1.7e− and the 1π → d donation by ~2.9e− evidence that one CO 1π bond has been cleaved to release one C 2p empty orbital. One neighboring O2− attacks this C 2p empty orbital to energetically favorably form one new C-O π bond, during which the 5σ → d donation41 by ~1.3e− more occurs at TS1 to subsequently form one new C-O σ bond (Fig. 3d and Supplementary Fig. 16), and these two new bonds constitute one C = O bond of CO2. As a result, the energy barrier at TS1 originates mainly from the depletion of CO 5σ electrons in order to form the new C-O σ bond. Further d → 2π* back-donation by ~0.8e− at TS1, which partially offsets the incremented positive charge of the adsorbed CO molecule (Fig. 3c and Supplementary Table 2), is conducive to the depletion of 5σ electrons. Although the total amount of the charges transferred from the CO 1π, 5σ, and 4σ orbitals onto the active site at the adsorption state is as large as ~4.1e− (Supplementary Table 2), these electronic charges are rapidly neutralized on this positively charged metallic particle, thus allowing the active site to further successively extract CO 5σ electrons only with a small energy input till the new C-O σ bond forms (Fig. 3b).
Besides, we calculate the average charge of core atoms in metallic Rh particle (Supplementary Table 3). The atoms maintain metallic states during the reaction cycle and have different electronic charges (negatively or positively charged) as a response of the different oxidation-reduction steps on the active sites, which can provide a channel for the electron transfer42. Meanwhile, the Rh atoms at nanofacets are covered by O species and these atoms stay in oxidation states (Supplementary Table 3) which should have no influence on the electronic communication during the CO oxidation on Rhmet/Al2O3. Thus, the electrons can be transferred between different vertex Rh sites of Rhmet/Al2O3 during the oxidation-reduction reaction cycle and the electronic communications couple two half-reactions to reduce the energy barrier.
Universal electronic communications on individual metallic particles
The electronic communications are not unique for metallic Rh particles, and we select the typical top metal (Pt) of a Sabatier volcano plot43 to study the influence of the electronic communications on activity. According to the particle sizes ( > 1.5 nm) required for the nonmetallic-to-metallic state transition6,20, we support metallic Pt particles or nonmetallic Pt clusters on Al2O3 to achieve Ptmet/Al2O3 or Ptnon/Al2O3, respectively (Supplementary Figs. 1 and 17–19). An average size for the Pt particles on Ptmet/Al2O3 or Ptnon/Al2O3 is approximately 2.4 or 0.7 nm, respectively (Fig. 4a or 4b). Ptmet/Al2O3 shows the metallic states evidenced by the XANES spectra at Pt L3-edge44 and the CO DRIFT spectra45,46 (Supplementary Figs. 20–21). As expected, Ptmet/Al2O3 exhibits much higher activity and lower Ea than Ptnon/Al2O3 in CO oxidation (Fig. 4c and inset, Supplementary Figs. 22–23), which suggests that metallic particles can lower the Ea via inter-site electronic communications. Such an inter-site electronic communication on metallic particles appears to be a common phenomenon in other catalytic reactions4,11,38,39,47,48,49,50.
a, b STEM images of Ptmet/Al2O3 (a) and Ptnon/Al2O3 (b) as well as the size distribution curves of the Pt particles (insets). c CO conversions as a function of temperature together with the extracted Ea (inset) over Ptmet/Al2O3 and Ptnon/Al2O3. d Schematic representation of two-half reactions coupling between two different active sites on one individual metallic particle or of the reaction without coupling occurring on isolated monoatomic active sites in CO oxidation. e, f STEM images of Rh1/Al2O3 (e) and Pt1/Al2O3 (f) and the atom-atom distance statistics (insets). g Ea values and TOFs of CO oxidation over Rh1/Al2O3 and Pt1/Al2O3 at 160 oC, which are obtained from Supplementary Figs. 29 and 30, respectively.
Analogous to two coupled electrochemical half-reactions, electronic communications on individual metallic particles can directly couple two half-reactions occurring on two different active sites in a spatially synchronized way in thermochemical catalytic oxidation4. As exemplified by using metallic Rh particle in CO oxidation (the upper half of Fig. 4d), upon being exposed in the reaction atmosphere, metallic Rh particles are often covered with dissociated oxygen via one half reaction: ① ½O2 + 2e− → O2−, which extracts two electrons from the active site I, inducing a charge redistribution on metallic Rh particle. As predicted by DFT calculations (Fig. 3), the charge redistribution energetically favorably triggers the other half reaction occurring on one entangled active site II: CO + O2− → CO2 + 2e−, thus realizing electrons delivery from active site II to active site I. Likewise, electrons will transfer from active site I to active site II, as each active site finishes the other half-reaction to close a catalytic cycle. Similar findings were also reported in heterogeneous catalysis11 or biocatalysis51, in which the coupling of two half-reactions could accelerate reaction rates to a great extent.
To investigate the driving force of two half-reactions coupling, we shut off channels for the inter-site electronic communications via separating single Rh or Pt atoms by an average distance of more than 3 nm in between on Al2O3 surfaces (the down half of Fig. 4d), as shown in the STEM images of Rh1/Al2O3 and Pt1/Al2O3 together with the insets of the atom-atom distance statistics in Fig. 4e or f, respectively (see also Supplementary Figs. 1 and 24–27). The relatively far distance between individual metal monoatoms (active sites) anchoring on the electronically insulating Al2O3 surfaces keeps the catalysts nonmetallic (Supplementary Figs. 20 and 28) and hampers two half-reactions coupling in CO oxidation. The Ea values and the TOFs (Fig. 4g, and Supplementary Figs. 29, 30) for Rh1/Al2O3 and Pt1/Al2O3 are in stark contrast with the values for Rhmet/Al2O3 and Ptmet/Al2O3, as would be expected for CO oxidation catalyzed by Rhnon/Al2O3 (Fig. 2b) or Ptnon/Al2O3 (Fig. 4c) without inter-site electronic communications. Hence, the electronic communications are the driving force to couple two half-reactions occurring on different active sites, thus lowering the energy barriers. This agrees well with the theoretical simulations (Fig. 3).
The driving force for these processes often originates from the electron-hole excitation, which leads to the generation of so-called “hot” electrons largely via chemisorption52. The generation probability and the number of the hot electrons are closely associated with the adsorption energy, and a large adsorption energy is often favorable to generate the large number of the hot electrons, leading to the increasing probability of inter-site electronic communications. As shown in Fig. 3a and Supplementary Fig. 7d, the adsorption energies ( > 1 eV) for CO and O are large enough to generate the reaction-required “hot” electrons1, thereby drving a smooth cycle by the inter-site electronic communications in CO oxidation.
Roles of metallic states in heterogeneous catalysis
As evidenced above, the metallic states of the particles are an essential prerequisite to realize the inter-site electronic communications with which to couple the two half-reactions. To explore the roles of the metallic states in heterogeneous catalysis, we scrutinize reported Ea values of CO oxidation over various supported metal particles with different sizes, as displayed in Fig. 5a (see also Supplementary Table 4 for more details). As particle sizes increases up to ~2 nm, an energy gap between the unoccupied states and occupied states becomes small, and ultimately a continuous energy level forms where the metallic states appear6,18,19,20. Correspondingly, the Ea value decreases from ~100 kJ mol−1 to a minimal value (~30 kJ mol−1), and almost remains constant (~0.3 eV) as the size further increases.
a Reported Ea values in CO oxidation over supported metal particle catalysts with different particle sizes. Gold pentagrams represent the catalysts prepared in this paper. Reation condition for the catalysts in this paper: 100 mg catalyst (40 ~ 60 mesh) was used; the total flow rate was 50 mL min−1, including 1.0 vol% CO, 21.0 vol% O2 and N2 as balance gas. The red hollow circles represent data from the literature. Typically, the black hollow triangles represent the Ea values in CO oxidation over single-atom catalysts with reducible supports. All the data for the reported catalysts and the related reaction conditions were in detail listed in Supplementary Table 4. b Schematic illustration of the energy level between HOMO (or LUMO) of adsorbed CO and O2 molecules (*CO and *O2) and LUBO (or HOBO) of nonmetallic particles or metallic particles.
To shed light on why the appearance of the metallic states can lower Ea in CO oxidation, we employ a combination of the frontier band orbital model53 and the interfacial energy alignment model31,32 to describe the energy level between molecular orbitals of adsorbed species and band orbitals of active sites. In Fig. 5b, according to the interfacial energy alignment model31,32, an adsorbed molecule’s HOMO can become pinned to the EF level of catalysts (which is also consistent with our DFT calculation). As for the nonmetallic particles with an energy gap between unoccupied states and occupied states, charge exchange between adsorbed species and catalysts requires surmounting energy barriers between HOMO and the lowest unoccupied band orbital (LUBO) or between the lowest unoccupied molecular orbital (LUMO) and the highest occupied band orbital (HOBO). As the metallic states form, HOBO and LUBO are merged on EF, thus lowering the energy barriers required for the charge exchange31,53.
The experimental evidence and the related discussion above clearly elucidate the roles of metallic states in lowering Ea in catalytic oxidation: (i) one is to allow HOBO and LUBO to merge on EF so as to eliminate energy gaps present for nonmetallic catalysts53, thus lowering energy barriers for charge exchanges between HOMO (or LUMO) of adsorbed species and LUBO (or HOBO) of active sites; (ii) the other is to give access to electronic communications for active sites on individual metallic particles as an electron transport, thus coupling two half-reactions occurring on two different active sites in a spatially synchronized way4, which lowers energy barriers for charge transfer between reactants. Hence, from the charge transfer view of point, the appearance of the metallic states can accelerate the reactant-catalyst charge exchanges and the catalyst-mediated reactant-reactant charge delivery, which makes CO oxidation follow preferentially a low-Ea electronic-communication-driven half-reactions coupling EDM11.
Discussion
The electron excitation and transfer play a critical role in heterogeneous catalysis. Typical pioneering work includes the electronic excitations by surface reactions, or so-called “exoelectron emission”54, and internal electron excitation within a catalysis system, or “chemicurrents”55. In particular, Somorjai’s group directly detected chemicurrents in catalytic reactions and found that chemicurrents were well correlated with the turnover rates, reflecting the prevalence of the electron transfer across a wide range of heterogenous reactions52. An important work from Ertl’s group is that excited electrons triggered catalytic reactions of CO oxidation with an Ea much smaller than that by phonon excitation56, implying that excited electrons by chemi- or physisorption of reactant’s molecule2,3,4,38 on one active site can be transferred to another entangled active site in reactions upon eligible electron conducting. In this work, we evidenced the presence of the electronic communications between active sites on individual Rh and Pt metallic nanoparticles in catalytic reaction of CO oxidation, and such an electronic communication might be commonly present in heterogeneous catalysis2,38, which holds for metallic materials such as Au, Pd, Ir, and so forth, catalyzing not only CO oxidation (Fig. 5a), but also for a broad spectrum of catalytic reactions, such as VOCs oxidation reaction49,57, three-way catalytic reaction50, and water-gas shift reaction45. Meanwhile, metal nanoparticles are required to bear the metallic states in order to initiate effective electron communications in reactions, and thus the metallic particles with a size as small as possible to expose more metal atoms on surfaces as active sites are favorable for activity enhancement. For further design, after realizing the metallic electronic structures of the metal catalysts, the atomic utilization and green economy of the catalysts may be important which need to be considered in order to achieve the goals of green chemistry. That is, the metal with sizes between 2 ~ 3 nm may be the optimal option but of course there may be some differences between reactions, which should be further investigated in the following work. Hence, this work provides a general strategy to rationally design and effectively develop improved metallic nanomaterials in heterogeneous catalysis.
These results give a unified explanation to disparate and controversial results in activity of single-atom catalysts and metallic particle catalysts. Owing to a lack of inter-site electronic communications, electron transfer for single-atom catalysts with electronically insulating supports is fulfilled merely via single active metal atoms by changing their valence states, which often requires surmounting a considerably high energy barrier to realize the oxidation ↔ reduction state transitions, thus leading to a relatively large Ea. Owing to inter-site electronic communications, metallic particle catalysts often show much higher activity than this kind of single-atom catalysts under identical conditions47,48,58. However, if suitable reducible supports selected can generate so strong metal-support interactions that only a low energy barrier is required to surmount for the electron transfer between single atoms and supports23,59, Ea in reactions over this kind of single-atom catalysts could be as low as that over metallic particle catalysts (Fig. 5a), in which case single-atom catalysts could have higher catalytic activity than metallic particles due to more exposed metal atoms as active sites60,61. Our results could also give a good interpretation for the so-called nanosized effects62. As the particle size increases, the appearance of metallic states enables the charge exchange energies to decrease down to a minimum, where a low-Ea-kinetic reaction coupling pathway is triggered via inter-site electronic communications, so catalytic activity increases up to a maximum. As the particle size further increases, catalytic activity often decreases owing to the decreasing number of surface metal atoms acting as active sites. Therefore, this work also has implications for an in-depth understanding of reaction mechanisms.
Methods
Preparation of Rhmet/Al2O3, Rhnon/Al2O3 and Rh1/Al2O3
The commercial non-reducible γ-Al2O3 (average particle size: 10-20 nm, 99.99%, 3AMaterials, China) was calcined at 550 °C for 6 h and then used as the support. The metallic Rhmet/Al2O3 containing metallic Rh nanoparticles was synthesized according to the colloid deposition method24. First, an ethylene glycol (EG) solution of NaOH (50 mL, 0.1 mol/L) was added into the EG solution of Rh(NO3)3 (50 mL, 5.8 mmol/L) with stirring at 140 °C for 3 h under the protection of Ar atmosphere to obtain Rh particle colloid solution. Then, 1 g γ-Al2O3 support was added into 16 mL Rh particle colloid solution under stirring at 80 °C for 3 h, followed by centrifugation with deionized water to remove EG, Na+ and NO3− and dried overnight. Finally, the obtained powder was calcined at 200 °C for 2 h in air. The nonmetallic Rhnon/Al2O3 was synthesized by the impregnation method. 1 g γ-Al2O3 support was dispersed in deionized water, and 1.91 mL of the (NH4)2RhCl4 solution (2.62 gRh/L) was added dropwise. Then the mixed solution was evaporated to dryness at 80 °C. Finally, the powder was dried and calcined at 500 °C for 3 h in air. The theoretical loading of Rh is 0.5 wt.%. Rh1/Al2O3 was prepared as the same process as Rhnon/Al2O3 except that the loading of Rh is 0.1 wt.%. The actual loadings of Rh were confirmed by inductively coupled plasma-atomic emission spectroscopy (ICP-AES).
Preparation of Ptmet/Al2O3, Ptnon/Al2O3 and Pt1/Al2O3
Ptmet/Al2O3 was also synthesized by the colloid deposition method. The EG solution of NaOH (50 mL, 0.05 mol/L) was added into an EG solution of H2PtCl6 (50 mL, 3.7 mmol/L) with stirring at 140 °C for 3 h under the protection of Ar atmosphere to obtain Pt particle colloid solution. Then, 1 g γ-Al2O3 support was added into 13 mL of the prepared Pt particle colloid solution under stirring at 80 °C for 3 h, followed by centrifugation with deionized water to remove EG, Na+ and Cl−. Finally, the obtained powders were calcined at 200 °C for 2 h in air. Ptnon/Al2O3 and Pt1/Al2O3 were synthesized by the impregnation method. 1 g γ-Al2O3 support was dispersed in deionized water, and 1.38 and 0.28 mL of the H2PtCl6 solution (3.6 gPt/L) was added dropwise to prepare Ptnon/Al2O3 and Pt1/Al2O3 respectively. Then the mixed solution was evaporated to dryness at 80 °C. Finally, the powders were dried and calcined at 400 °C for 2 h in air. The actual loadings of Pt were measured by ICP-AES.
X-ray diffraction (XRD) patterns
XRD patterns were collected on a D8 advance X-ray diffractometer (Bruker/AXS, Germany) at 40 kV and 40 mA using monochromatized Cu Kα radiation (λ = 1.5405 Å).
Transmission electron microscopy (TEM) and aberration-corrected high-angle annular dark-field scanning transmission electron microscopy (AC-STEM)
TEM experiments were carried out with a JEOL JEM-2100F field-emission gun transmission electron microscope operated at an accelerating voltage of 200 kV and equipped with an ultra-high-resolution pole-piece that provides a point-resolution better than 0.19 nm. Fine powders of the materials were dispersed in ethanol, sonicated, and sprayed on a carbon coated copper grid, and then allowed to air-dry. AC-STEM experiments and energy dispersive X-ray spectroscopy (EDS) elemental mapping were carried out in an aberration-corrected scanning/transmission electron microscope (ThermoFisher Thermis Z) equipped with SuperEDX detector at an accelerating voltage of 300 kV. An aberration-corrected scanning/transmission electron microscope (Hitachi HF5000) was also used for microscopic analysis.
X-ray absorption spectroscopy (XAS)
XAS includes X-ray absorption near-edge structure (XANES) spectra and extended X-ray absorption fine structure (EXAFS) spectra was obtained at beamline BL14W1 of the Shanghai Synchrotron Radiation Facility (SSRF). The data at Rh K-edge and Pt L3-edge were collected with a fixed exit monochromator using Si(311) and Si(111) crystals, respectively.
X-ray photoelectron spectra (XPS)
XPS were recorded on a Kratos Axis Ultra-DLD system with a charge neutralizer and a 150 W Al (Mono) X-ray gun (1486.6 eV) with a delay-line detector (DLD). Spectra were acquired at normal emission with a passing energy of 40 eV. Charging effects were corrected by adjusting binding energy of C 1 s to 284.8 eV.
Catalytic evaluation
All the samples were pretreated at 300 °C for 30 min in Ar before experiments. CO oxidation was carried out in a fixed-bed quartz reactor. In each activity test, 100 mg catalyst (40 ~ 60 mesh) was used. The total flow rate was 50 mL min−1, including 1.0 vol% CO, 21.0 vol% O2 and N2 as balance gas. The CO and CO2 concentrations were detected by GC7890B (Agilent, USA) with FID and TCD detectors. When determining the reaction TOF, 5 mg Rhmet/Al2O3 and 200 mg Rhnon/Al2O3 were used and Si3N4 (α-phase, 40 ~ 60 mesh, Aldrich) was blended as diluents to keep the gas hourly space velocity (GHSV) consistent.
Temperature-programmed desorption (TPD)
O2-TPD or CO-TPD profiles were obtained by AutoChem II 2920 (Micromeritics, USA). O2-TPD experiments were performed in a U-shaped quartz reactor with a TCD as detector. In each test,100 mg sample was pretreated with He gas (50 mL min−1) at 250 °C for 1 h. After cooling down to room temperature, the catalyst was treated with O2 gas (30 mL min−1) or 5% CO/He (30 mL min−1) for 1 h for O2-TPD or CO-TPD, respectively. After the adsorption was completed, pure He gas (50 mL min−1) was introduced to remove the weak physical adsorption O2 or CO on the surface. Finally, the sample was then heated to 300 °C at a rate of 10 °C min−1.
Diffuse-reflectance infrared Fourier-transform (DRIFT) spectra
DRIFT spectra were conducted by accumulating 64 scans at 4 cm−1 resolution from 4000 to 1000 cm−1 on a Nicolet iS50 FTIR spectrometer equipped with a Harrick Scientific DRIFT cell and a mercury-cadmium-telluride MCT/A detector.
For ex situ measurement, samples were pretreated at 300 °C for 1 h in a flow of 50 mL min−1 He to remove adsorbed water and carbon dioxide and then cooled to 30 °C under He flow to obtain a background spectrum which should be deducted from the spectra of samples. Next, 25 mL min−1 1% CO/He mixed gas was introduced. After the CO saturation, 50 mL min−1 He was used to remove the gas phase CO from cell. Finally, the spectra were obtained at 30 °C. The spectra at 150 oC in Fig. 2e and Supplementary Fig. 12 were collected in the O2/He mixed gas after the temperature increased to 150 oC from 30 oC.
For temperature-programmed in situ DRIFT spectra, samples were pretreated at 300 °C for 1 h in a flow of 50 mL min−1 He to remove adsorbed water and carbon dioxide and then cooled to 30 °C under He flow. Next, samples were exposed to a flow of 25 mL min−1 mixed gas (1% CO + 20% O2 in He) and heated to 300 °C at a temperature rate of 5 °C min−1. The experiments were also performed with Al2O3 to subtract the signal of gaseous CO from the spectra.
DFT calculation
Density functional theory (DFT) calculations were performed using VASP 6.2.1 packages63 with projected augmented wave (PAW) pseudo-potentials64,65. The exchange-correlation energy was treated based on the generalized gradient approximation (GGA) by using Perdew–Burke–Ernzerhof (PBE) functional66. The plane-wave cutoff energy was set to 400 eV. The DFT-D3(BJ) method of Grimme67,68 was employed to describe long-range VDW interactions. The Monkhorst–Pack scheme with a k-point separation length of 0.05 Å−1 was utilized for sampling the first Brillion zone69. To correct the zero-point energy for the reaction barrier, the vibrational frequency calculations were performed via the finite-difference approach. All atoms were fully relaxed in the calculations. The Quasi-Newton l-BFGS method was used for geometry relaxation until the maximal force on each degree of freedom was less than 0.05 eV Å−1. To derive the free energy reaction profiles, we followed the same approach as our previous work70. The standard thermodynamic data71 were utilized to obtain the temperature and pressure contributions.
Data availability
Source data are provided as a Source Data file. All other data of this study are available from the corresponding author upon reasonable request.
References
Gergen, B., Nienhaus, H., Weinberg, W. H. & McFarland, E. W. Chemically induced electronic excitations at metal surfaces. Science 294, 2521–2523 (2001).
Novo, C., Funston, A. M. & Mulvaney, P. Direct observation of chemical reactions on single gold nanocrystals using surface plasmon spectroscopy. Nat. Nanotechnol. 3, 598–602 (2008).
Adams, J. S., Kromer, M. L., Rodríguez-López, J. & Flaherty, D. W. Unifying concepts in electro- and thermocatalysis toward hydrogen peroxide production. J. Am. Chem. Soc. 143, 7940–7957 (2021).
Ryu, J. et al. Thermochemical aerobic oxidation catalysis in water can be analysed as two coupled electrochemical half-reactions. Nat. Catal. 4, 742–752 (2021).
Gates, B. C. Individual nanoparticles in action. Nat. Nanotechnol. 3, 583–584 (2008).
Valden, M., Lai, X. & Goodman, D. W. Onset of catalytic activity of gold clusters on titania with the appearance of nonmetallic properties. Science 281, 1647–1650 (1998).
Fu, Q., Saltsburg, H. & Flytzani-Stephanopoulos, M. Active nonmetallic Au and Pt species on ceria-based water-gas shift catalysts. Science 301, 935–938 (2003).
Xu, Z. et al. Size-dependent catalytic activity of supported metal clusters. Nature 372, 346–348 (1994).
Kaden, W. E., Wu, T., Kunkel, W. A. & Anderson, S. L. Electronic structure controls reactivity of size-selected Pd clusters adsorbed on TiO2 surfaces. Science 326, 826–829 (2009).
Maurer, F. et al. Tracking the formation, fate and consequence for catalytic activity of Pt single sites on CeO2. Nat. Catal. 3, 824–833 (2020).
Huang, Z. et al. Interplay between remote single-atom active sites triggers speedy catalytic oxidation. Chem. 8, 3008–3017 (2022).
Freund, H.-J., Meijer, G., Scheffler, M., Schlögl, R. & Wolf, M. CO oxidation as a prototypical reaction for heterogeneous processes. Angew. Chem. Int. Ed. 50, 10064–10094 (2011).
Sabatier, P. et al. La Catalyse En Chimie Organique; Librairie Polytechnique, Paris et Liége: Paris. (1920).
Hoffmann, R. A chemical and theoretical way to look at bonding on surfaces. Rev. Mod. Phys. 60, 601 (1988).
Nolte, P. et al. Shape changes of supported Rh nanoparticles during oxidation and reduction cycles. Science 321, 1654–1658 (2008).
Falsig, H. et al. Trends in the catalytic CO oxidation activity of nanoparticles. Angew. Chem. Int. Ed. 47, 4835–4839 (2008).
Kubo, R. Electronic properties of metallic fine particles. I. J. Phys. Soc. Jpn. 17, 975–986 (1962).
Baetzold, R. C. Electronic properties of metal clusters: size effects. Inorg. Chem. 20, 118–123 (1981).
Li, L. et al. Investigation of catalytic finite-size-effects of platinum metal clusters. J. Phys. Chem. Lett. 4, 222–226 (2013).
Xu, C., Lai, X., Zajac, G. W. & Goodman, D. W. Scanning tunneling microscopy studies of the TiO2(110) surface: structure and the nucleation growth of Pd. Phys. Rev. B 56, 13464–13482 (1997).
Hu, S. & Li, W.-X. Sabatier principle of metal-support interaction for design of ultrastable metal nanocatalysts. Science 374, 1360–1365 (2021).
Bruix, A. et al. A new type of strong metal–support interaction and the production of H2 through the transformation of water on Pt/CeO2(111) and Pt/CeOx/TiO2(110) catalysts. J. Am. Chem. Soc. 134, 8968–8974 (2012).
Hu, P. et al. Electronic metal–support interactions in single-atom catalysts. Angew. Chem. Int. Ed. 53, 3418–3421 (2014).
Wang, Y., Ren, J., Deng, K., Gui, L. & Tang, Y. Preparation of tractable platinum, rhodium, and ruthenium nanoclusters with small particle size in organic media. Chem. Mater. 12, 1622–1627 (2000).
Smith, N. V., Wertheim, G. K., Hüfner, S. & Traum, M. M. Photoemission spectra and band structures of d-band metals. IV. X-ray photoemission spectra and densities of states in Rh, Pd, Ag, Ir, Pt, and Au. Phys. Rev. B 10, 3197–3206 (1974).
Filatova, E. O. & Konashuk, A. S. Interpretation of the changing the band gap of Al2O3 depending on its crystalline form: connection with different local symmetries. J. Phys. Chem. C. 119, 20755–20761 (2015).
Fu, Q., Wagner, T., Olliges, S. & Carstanjen, H.-D. Metal−oxide interfacial reactions: encapsulation of Pd on TiO2(110). J. Phys. Chem. B 109, 944–951 (2005).
Weast, R. C., Astle, M. J., Ed. CRC Handbook of Chemistry and Physics, 63rd ed. (CRC Press, 1982).
Hülsey, M. J. et al. In situ spectroscopy-guided engineering of rhodium single-atom catalysts for CO oxidation. Nat. Commun. 10, 1330 (2019).
Bunluesin, T., Putna, E. S. & Gorte, R. J. A comparison of CO oxidation on ceria-supported Pt, Pd, and Rh. Catal. Lett. 41, 1–5 (1996).
Hill, I. G., Rajagopal, A., Kahn, A. & Hu, Y. Molecular level alignment at organic semiconductor-metal interfaces. Appl. Phys. Lett. 73, 662–664 (1998).
Greiner, M. T. et al. Universal energy-level alignment of molecules on metal oxides. Nat. Mater. 11, 76–81 (2012).
Wovchko, E. A. & Yates, J. T. Activation of O2 on a photochemically generated RhI site on an Al2O3 surface: low-temperature O2 dissociation and CO oxidation. J. Am. Chem. Soc. 120, 10523–10527 (1998).
Cai, J. et al. Formation of different Rh–O species on Rh(110) and their reaction with CO. ACS Catal. 13, 11–18 (2023).
Wintterlin, J., Völkening, S., Janssens, T. V. W., Zambelli, T. & Ertl, G. Atomic and macroscopic reaction rates of a surface-catalyzed reaction. Science 278, 1931–1934 (1997).
Stenlid, J. H. & Brinck, T. Extending the σ-hole concept to metals: an electrostatic interpretation of the effects of nanostructure in gold and platinum catalysis. J. Am. Chem. Soc. 139, 11012–11015 (2017).
Yoon, B. et al. Charging effects on bonding and catalyzed oxidation of CO on Au8 clusters on MgO. Science 307, 403–407 (2005).
Zou, N. et al. Cooperative communication within and between single nanocatalysts. Nat. Chem. 10, 607–614 (2018).
Suchorski, Y. et al. Resolving multifrequential oscillations and nanoscale interfacet communication in single-particle catalysis. Science 372, 1314–1318 (2021).
Anderson, A. B. Molecular orbital study of the interaction of carbon monoxide and carbon dioxide with copper (100). Surf. Sci. 62, 119–132 (1977).
Doyen, G. & Ertl, G. Theory of carbon monoxide chemisorption on transition metals. Surf. Sci. 43, 197–229 (1974).
Hervier, A., Renzas, J. R., Park, J. Y. & Somorjai, G. A. Hydrogen oxidation-driven hot electron flow detected by catalytic nanodiodes. Nano Lett. 9, 3930–3933 (2009).
Nørskov, J. K. et al. Origin of the overpotential for oxygen reduction at a fuel-cell cathode. J. Phys. Chem. B 108, 17886–17892 (2004).
Bearden, J. A. & Burr, A. F. Reevaluation of X-ray atomic energy levels. Rev. Mod. Phys. 39, 125–142 (1967).
Ding, K. et al. Identification of active sites in CO oxidation and water-gas shift over supported Pt catalysts. Science 350, 189–192 (2015).
Niu, B. et al. Modulating the electronic states of Pt nanoparticles on reducible metal–organic frameworks for boosting the oxidation of volatile organic compounds. Environ. Sci. Technol. 58, 4428–4437 (2024).
Guo, L.-W. et al. Contributions of distinct gold species to catalytic reactivity for carbon monoxide oxidation. Nat. Commun. 7, 13481 (2016).
Dessal, C. et al. Dynamics of single Pt atoms on alumina during CO oxidation monitored by operando X-ray and infrared spectroscopies. ACS Catal. 9, 5752–5759 (2019).
Goodman, E. D. et al. Catalyst deactivation via decomposition into single atoms and the role of metal loading. Nat. Catal. 2, 748–755 (2019).
Jeong, H. et al. Highly durable metal ensemble catalysts with full dispersion for automotive applications beyond single-atom catalysts. Nat. Catal. 3, 368–375 (2020).
Reeve, H. A. et al. Enzymes as modular catalysts for redox half-reactions in H2-powered chemical synthesis: from biology to technology. Biochem. J. 474, 215–230 (2017).
Park, J. Y., Baker, L. R. & Somorjai, G. A. Role of hot electrons and metal–oxide interfaces in surface chemistry and catalytic reactions. Chem. Rev. 115, 2781–2817 (2015).
Xu, D. et al. An activity descriptor for perovskite oxides in catalysis. Chem. Catal. 2, 1163–1176 (2022).
Kramer, J. Spitzenzähler und zählrohr bei metallographischen oberflächenuntersuchungen. Z. Phys. 125, 739–756 (1949).
Nienhaus, H. et al. Electron-hole pair creation at Ag and Cu surfaces by adsorption of atomic hydrogen and deuterium. Phys. Rev. Lett. 82, 446–449 (1999).
Bonn, M. et al. Phonon- versus electron-mediated desorption and oxidation of CO on Ru(0001). Science 285, 1042–1045 (1999).
Liu, J. et al. Benzene abatement catalyzed by ceria-supported platinum nanoparticles and single atoms. Chem. Eng. J. 467, 143407 (2023).
Wang, J., Tan, H., Yu, S. & Zhou, K. Morphological effects of gold clusters on the reactivity of ceria surface oxygen. ACS Catal. 5, 2873–2881 (2015).
Wang, L. et al. Supported rhodium catalysts for ammonia–borane hydrolysis: dependence of the catalytic activity on the highest occupied state of the single rhodium atoms. Angew. Chem. Int. Ed. 56, 4712–4718 (2017).
Qiao, B. et al. Single-atom catalysis of CO oxidation using Pt1/FeOx. Nat. Chem. 3, 634–641 (2011).
DeRita, L. et al. Catalyst architecture for stable single atom dispersion enables site-specific spectroscopic and reactivity measurements of CO adsorbed to Pt atoms, oxidized Pt clusters, and metallic Pt clusters on TiO2. J. Am. Chem. Soc. 139, 14150–14165 (2017).
Liu, L. & Corma, A. Metal catalysts for heterogeneous catalysis: from single atoms to nanoclusters and nanoparticles. Chem. Rev. 118, 4981–5079 (2018).
Kresse, G. & Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 6, 15–50 (1996).
Kresse, G. & Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 59, 1758–1775 (1999).
Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 50, 17953–17979 (1994).
Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996).
Grimme, S., Antony, J., Ehrlich, S. & Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 132, 154104 (2010).
Grimme, S., Ehrlich, S. & Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 32, 1456–1465 (2011).
Monkhorst, H. J. & Pack, J. D. Special points for brillouin-zone integrations. Phys. Rev. B 13, 5188–5192 (1976).
Wei, G.-F., Shang, C. & Liu, Z.-P. Confined platinum nanoparticle in carbon nanotube: structure and oxidation. Phys. Chem. Chem. Phys. 17, 2078–2087 (2015).
Lide, D. R. CRC Handbook Of Chemistry And Physics. (CRC press, 2004).
Acknowledgements
This work is supported by National Key R&D Program of China 2022YFB3504100, 2021YFB3500600 (to X.T.), National Natural Science Foundation of China 22276036, 22276037, 22173069, 22072090, 22272106 (to X.T., Z.M., G.W., X.L., and X.L.) and Shanghai Municipal Natural Science Foundation 21ZR1467800 (to G.W.). XAS at Rh K-edge and Pt L3-edge were performed at beamline BL14W1 of the Shanghai Synchrotron Radiation Facility (SSRF).
Author information
Authors and Affiliations
Contributions
X.T. directed the project; D.X., X.F., C.X., and J.C. synthesized the samples, designed and conducted the experiments; Y.J. and G.W. did the theoretical calculation work; B.H., G.C., L.C., and X.L. provided equipment for electron microscopy experiments; X.T., Y.C., D.X., and W.Q. analyzed the data and interpreted the results. D.X., X.T., Z.M., and Y.C. wrote the manuscript. All authors have given approval to the final version of the manuscript.
Corresponding authors
Ethics declarations
Competing interests
The authors declare no competing interests.
Peer review
Peer review information
Nature Communications thanks Zhimin Ao and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Source data
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/.
About this article
Cite this article
Xu, D., Jin, Y., He, B. et al. Electronic communications between active sites on individual metallic nanoparticles in catalysis. Nat Commun 15, 8614 (2024). https://doi.org/10.1038/s41467-024-52997-w
Received:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41467-024-52997-w