Shouyao Hua,
Jiaxin Gonga,
Yu Taoa,
Runze Maa,
Jianping Guana,
Xu Liua,
Jinhua Hua,
Jun Yan*a,
Shibin Wangb,
Zedong Zhangc,
Xiao Liangc,
Zechao Zhuangcd,
Yunhu Han
e,
Xusheng Zheng
f,
Wensheng Yan
f,
Chengjin Cheng,
Wei Zhu
g,
Dingsheng Wang
c and
Yu Xiong
*a
aDepartment of Chemistry and Chemical Engineering, Central South University, Changsha 410083, China. E-mail: yanjun@csu.edu.cn; thomas153@126.com
bCollege of Chemical Engineering, Zhejiang University of Technology, Hangzhou 310032, China
cDepartment of Chemistry, Tsinghua University, Beijing 100084, China
dDepartment of Chemical Engineering, Columbia University, New York, NY 10027, USA
eFrontiers Science Center for Flexible Electronics, Xi'an Institute of Flexible Electronics (IFE) and Xi'an Institute of Biomedical Materials & Engineering, Northwestern Polytechnical University, Xi'an 710072, China
fNational Synchrotron Radiation Laboratory, University of Science and Technology of China, Hefei 230029, China
gState Key Lab of Organic-Inorganic Composites, Beijing Advanced Innovation Center for Soft Matter Science and Engineering, Beijing University of Chemical Technology, Beijing 100029, China
First published on 27th January 2025
The simultaneous regulation of particle size, surface coordinated environment and composition for Pt-based intermetallic compound (Pt-IMC) nanoparticles to manipulate their reactivity for energy storage is of great importance. Herein, we report a general synthetic method for Pt-IMCs using SBA-15 for coordination-in-pipe engineering. The particle size can be regulated to 3–9 nm by carrying out the coordination in pipes with different diameters and the coordination number of the interface metal atoms can be adjusted by altering the N source. Moreover, this strategy can also be expanded to the synthesis of Pt-IMCs with the majority of fourth period transition metals (Ti, V, Cr, Mn, Fe, Co, Ni, Cu, Zn). The Pt3Co IMC using 1,10-phenanthroline as the nitrogen source (Pt3Co@CN) shows the highest catalytic performance in the methanol oxidation reaction (MOR; 2.19 A mgPt−1) among the investigated nitrogen sources. The high chemical states of surface Pt and Co, affected by the nitrogen coordination number at the angstrom scale, facilitate electron accumulation on active sites, reduce the activation energy of the rate-determining step and enhance the catalytic performance of Pt-IMCs in the MOR.
Developing synthetic strategies towards IMCs has attracted much attention in recent years. The solvothermal strategy efficiently adjusts the morphology via a capping agent-assisted crystallization process, while an obstacle to downsizing IMCs via solvothermal strategy is difficult to overcome, which means dissipating a large amount of inner Pt sites.34–37 Recently, several studies have been reported using a pyrolysis-based strategy,38–40 which can efficiently synthesize small sized IMCs (ca. 5 nm). For example, Yang et al.41 developed a general sulfur-anchoring strategy to obtain Pt-based IMCs with different elements, which show high catalytic performance in the acidic oxygen reduction reaction. Cheng and co-workers synthesized Pt-IMCs on ZIF-8-derived nitrogen-doped porous carbon by a wet-impregnation method, which exhibited high ORR performance.42 The unsaturated Pt or M (M = Fe Co Ni et al.) atoms on the surface are coordinated by elements with high electronegativity (C, N, O, S et al.) on the support of such catalysts,43–49 and thus the electrons will transfer from the Pt or M to the coordinated element, which effectively adjusts their catalytic performance, especially among small particles with a higher ratio of surface anchoring metal atoms. Additionally, owing to the discontinuous molecular orbitals in small nanoparticles, manipulating the particle size of IMCs is also vital to realizing their catalytic performance.50–53 The reported works concentrated more on the IMCs loaded on settled coordinated sites of the prefabricated support. For example, the synthesis of Pt-IMCs on an S-containing support results in smaller particle sizes (<5 nm), while using only N- or O-containing groups will remarkably increase the particle size,54 indicating that it is irresolvable to manipulate the particle size with the same coordinated environment and precisely adjust their catalytic performance by this strategy.
Due to their high surface area, thermostability, and porosity, zeolites are widely used as a hard template to synthesize porous carbon.41,55 Among those zeolites, SBA-15 possesses ordered pipes with adjustable diameter, which could be an ideal template to control the particle sizes of IMCs.56,57 The pipe-like mesopores can effectively hinder the growth and regulate the particle size of IMCs. Herein, we report a general synthetic method for Pt-based IMCs (PtM; M = Ti, V, Cr, Mn, Fe, Co, Ni, Cu, Zn) with controllable size and coordinated surface sites on a nitrogen-doped carbon catalyst using SBA-15 for coordination-in-pipe engineering. The particle sizes can be controlled by regulating the pipe diameter of SBA-15 at the nanoscale (3–9 nm). The chemical states of Pt and Co can also be manipulated by adjusting the coordination number with nitrogen at the angstrom scale, which is ascribed to the composition and structure of the N sources before pyrolysis. Moreover, the as-synthesized Pt3Co using 1,10-phenanthroline as the N source shows the highest catalytic performance in the methanol oxidation reaction (MOR) with a mass activity of 2.19 A mgPt−1, which is much higher than that of Pt3Co synthesized using other N sources. Density functional theory (DFT) combined with experimental results illustrate that the high chemical state of the metals effectively facilitates electron accumulation at the active sites, reducing the free energy of the intermediate states, and the activation energy of the rate-determining step (*CO–*COOH) in the MOR, which accelerates the reaction.
Pt3Co IMCs, synthesized using 1,10-phenanthroline and SBA-15 (average diameter: 8 nm), are taken as an example (namely, Pt3Co@CN) to illustrate the typical characterization of Pt-IMCs involved in this study. The broad peaks, especially the superlattice diffraction peak (located at 32.8°) in the powder X-ray diffraction (XRD) pattern, can be attributed to Pt3Co IMCs, which matches with the Joint Committee on Powder Diffraction Standards (JCPDS PDF#29-0499; Fig. 2a). From the transmission electron microscopy (TEM) images, we can observe that Pt3Co@CN maintained the pipe-like mesoporous features of the original SBA-15 (Fig. S2 and S3†). Numerous nanoparticles with an average size of 4.5 nm can be observed in the TEM images (Fig. 2b). From energy dispersive X-ray spectroscopy (EDS) mapping, Pt and Co are mainly detected in the nanoparticles, and N is homogeneously dispersed in the support (Fig. 2c). Owing to the different Z-contrast, Pt–Pt and Co–Pt–Co atomic columns can be clearly identified in the aberration-corrected high-angle annular dark-field scanning transmission electron microscopy images (AC-HAADF-STEM; Fig. 2d and e), which matches well with the corresponding crystal model (the inset of Fig. 2e). The corresponding fast Fourier transform (FFT) pattern shows that Pt3Co IMC is oriented along the [01] direction (the inset of Fig. 2d) with the lattice spacings measured to be 0.384 nm and 0.279 nm, which can be attributed to the (100) and (110) planes of Pt3Co (Pm
m), respectively.
To explore the effect of the pipe diameter on the size of the IMCs, SBA-15 samples with average diameters of 4 nm and 18 nm (characterization is shown in Fig. S4–S6†) were used (named Pt3Co (4) and Pt3Co (18), respectively). With increasing average pipe diameter, the half peak width in the XRD pattern becomes sharper, suggesting an increase in the IMCs size (Fig. 2a). The average particle sizes of Pt3Co (4) and Pt3Co (18) were 3.02 nm, and 8.24 nm, respectively (Fig. 2f and g) with Pt and Co concentrated in the nanoparticles and N distributed in the support (Fig. 2h and i). Moreover, the synthesis of Pt3Co without SBA-15 leads to phase separation (disordered PtCo alloy and Pt) and larger particle size (ca. 38 nm; Fig. S7†). When synthesizing without a N source, IMC nanoparticles will grow out of the pipes, resulting in larger nanoparticles (ca. 16 nm; Fig. S8†). These results illustrate that (1) SBA-15 inhibits both the growth of IMCs and the phase separation of Pt and Co; (2) the carbonization of 1,10-phenanthroline hinders the migration of Pt and Co along the pipes. Furthermore, this strategy can also work with other mesoporous zeolites. When hollow silica with spherical mesopores is used in the synthesis, Pt3Co IMCs with an average particle size of 7.8 nm (Pt3Co@mshs; Fig. S9†) are also successfully obtained.
To further explore the influence of the nitrogen source on the microstructure of Pt3Co IMCs, a series of characterizations for Pt3Co@CN, Pt3Co (quin) and Pt3Co (mim) are carried out. X-ray photoelectron spectroscopy (XPS) survey spectra show that C, N, Co, and Pt are the primary elements in Pt3Co (N sources) (Fig. S23†). Three deconvoluted peaks, attributed to graphitic C (284.8 eV), CC–N (286.2 eV) and C–C
N (288.4 eV), were observed in the C 1s spectra,58,59 which are almost identical with an average deviation of less than 1.0% (Table S5 and Fig. S24–S26†). The deconvoluted N 1s peaks at 398.6, 399.5, 400.5, and 401.8 eV can be ascribed to pyridinic, coordinated (Pt/Co–N), pyrrolic, and graphitic N, respectively60,61 (Fig. 3a). Surprisingly, though the microenvironment and the coordinated structure are different before pyrolysis, only negligible disparity with an average deviation of less than 3.7% in the ratio of N types of Pt3Co (N sources) can be seen in the N 1s XPS spectra (Table S5†). C K-edge near-edge X-ray absorption fine structure (NEXAFS) spectra (Fig. S27†) exhibit that the spilt resonances62 at 284.3, 286.4, and 288.9 eV can be attributed to the dipole transitions from the C 1s to π* orbitals of C
C, C–N and C
N, respectively. N K-edge NEXAFS spectra are shown in Fig. 3b; we can find spilt resonances63 located at 398.1 eV (pyridinic N), 399.0 eV (pyrrolic N), and 401.1 eV (graphitic N), respectively. Notably, the resonances of pyridinic N and pyrrolic N in Pt3Co@CN are broader than those in the other samples, indicating that more electrons transfer from Pt to N.64
The nitrogen contents of Pt3Co@CN, Pt3Co (quin) and Pt3Co (mim) were detected using a nitrogen analyser (1.69 wt%, 1.08 wt%, 0.83 wt%, respectively; Table S6†), and are lower than their theoretical nitrogen content (10.75 wt%, 7.67 wt%, 22.68 wt%, respectively). This result may be affected by the degree of decomposition of the N species (15.7%, 14.1%, and 3.6% retained, respectively). Pt and Co chelated coordinated with 1,10-phenanthroline show the highest thermostability, the single-dentate coordination with quinoline shows moderate thermal stability and the large content of excessive nitrogen without coordination in 2-methylimidazole is easier to decompose. Though near 0 chemical state of Pt are detected in the Pt 4f XPS spectra of Pt3Co (N sources), positive shifts of binding energies in the spectra from Pt3Co (mim) (0.32 eV), Pt3Co (quin) (0.40 eV), and Pt3Co@CN (0.44 eV) to Pt0 are also detected (Fig. 3c). The deconvoluted signals of Co2+ and Co0 species are detected in the Co 2p XPS spectra65 and the ratio of Co2+ in Pt3Co@CN is higher than that in Pt3Co (mim) and Pt3Co (quin) (Fig. 3d). Moreover, the increasing trend of the chemical state of Co and Pt is correlated with that of nitrogen content in Pt3Co (N sources). According to the results from the N 1s XPS and K-edge NEXAFS spectra combined with ICP-OES, the ratios of coordinated N are similar among the Pt3Co (N sources), but the highest nitrogen content leads to the transference of most electrons from the IMCs to nitrogen and thus the highest chemical state of Co and Pt in Pt3Co@CN. Additionally, to explore the effect of the metal content on the chemical states, we also synthesize Pt3Co with different metal dosages (the molar ratio of Pt and Co is not changed; namely, Pt3Co-x, where x represents the metal dosage; characterizations are shown in Fig. S28–S31†). The Pt 4f and Co 2p XPS peaks of Pt3Co-x do not show obvious shifts, proving that the metal content has little effect on the metal valence state (Fig. S32†). However, the impact of varying N sources on the particle size is relatively minor with a standard deviation of 7.6%, compared to the significant influence of different SBA-15 with a standard deviation of 38%.
To further explore the coordination parameters of Pt in Pt3Co (N sources), Pt L3-edge X-ray absorption fine structure (XAFS) analysis is performed. According to the X-ray absorption near-edge structure (XANES) spectra of Pt3Co (N sources) (Fig. 3e), the height of the white line peaks shows the same trends as the Pt 4f XPS spectra [Pt3Co@CN > Pt3Co(quin) > Pt3Co (min)] and slightly higher than that of Pt foil, further confirming that the electrons of surface Pt are transferred to N in the IMCs. The main peaks in the Fourier-transformed EXAFS of Pt3Co (mim), Pt3Co (quin) and Pt3Co@CN are located at 2.35 Å, 2.28 Å, and 2.19 Å, respectively, attributed to the co-existence of Pt–Co and Pt–Pt bonds, which are obviously lower than the Pt–Pt peak (2.45 Å) in Pt foil (Fig. 3f). The notable difference in Pt–Co may be ascribed to different strengths of metal-support interactions in the three samples. Though the signal intensities below 2.0 Å for Pt3Co (N sources) are all higher than that of Pt foil, which may be ascribed to the Pt–N bond, the relatively low ratio of Pt–N bond can hardly be totally separated from the metallic bond with high coordination number.
To evaluate the generality of this strategy, Pt–M IMCs (M = Ti, V, Cr, Mn, Fe, Ni, Cu and Zn) were prepared. XRD patterns showed that the diffraction peaks of Pt-IMCs matched the corresponding JCPDS standard cards well (Fig. S33–S40†), indicating the successful synthesis of a series of Pt–M IMCs. Small nanoparticles with average sizes ranging from 3.12 to 4.97 nm (Fig. 4a–h, S41–S48 and Table S7†) are distributed on the CN support densely without agglomeration in the TEM images, indicating that the growth of Pt–M IMCs can be inhibited from sintering using this strategy. HAADF-STEM images and corresponding EDS mappings (Fig. S49–S56†) illustrate that Pt and M were both uniformly distributed on nanoparticles. The atomic phase structure of IMCs corresponding to the atomic model crystal can be directly observed by AC-HAADF-STEM (Fig. 4i–p), confirming the high generality of this strategy. The measured atomic column spacing is shown in Table S8,† which is consistent with the theoretical values of the (110) and (100) crystal face spacing. Additionally, S sources, instead of N sources, were also applied in the synthesis of IMCs. Though the S sources can also hinder the growth of IMCs, the intensity of the superlattice diffraction peak becomes weak, indicating that S sources may not be suitable for this strategy (Fig. S57–S59†).
Pt3Co (quin) shows a good mass activity of 1.26 A mgPt−1 and Pt3Co (mim) exhibits a moderate mass activity of 0.48 A mgPt−1. Additionally, Pt3Co@CN shows the lowest onset potential (0.45 V) and Pt3Co (min) (0.63 V) is even more positively shifted than commercial Pt/C (0.58 V). These results suggest that the chemical states of Pt and Co, which are attributed to the different nitrogen contents in the catalysts, can highly affect the MOR performance (Fig. 5e). Moreover, the ratio of forward current and reverse current (If/Ib) of Pt3Co IMCs shows no obvious difference and is much higher than that of commercial Pt/C (Fig. 5f), indicating that the CO resistance is improved by the formation of IMCs.
To further investigate the catalytic mechanism of MOR catalyzed by Pt3Co@CN, in situ Fourier transform infrared (FTIR) analysis is performed (Fig. S63†). The peak with high absorbance, located at 1659 cm−1 in the potential range from 0.7–1.2 V, can be attributed to the HCOO*, indicating that HCOO* is the key intermediate in this reaction.66 Two positive peaks located at 2054 cm−1 and 2341 cm−1 can be ascribed to the linear absorption of CO and the formation of CO2 in the potential range from 0.6–1.2 V,67 suggesting that the poisoning of CO and the desorption of CO2 do not affect the reactivity of Pt3Co@CN.
To further explore the influence of the nitrogen content and coordination environment on the MOR performance, DFT calculations were subsequently carried out. Upon formation of the Pt–N bond in Pt3Co@CN, the heterointerfaces between the (110) facet of Pt3Co IMCs and support are fabricated. The surface Pt and Co atoms in Pt3Co are coordinated by the nitrogen of the support with different coordination numbers (Pt3Co–N8 and Pt3Co–N12; Fig. S64†). As shown in Fig. 5g, S65 and S66,† Pt3Co–N12 shows a lower global free energy profile in the reaction process than Pt3Co–N8. Moreover, the reaction energy for the elementary step of CO* hydrogenation to COOH* was calculated to be 0.69 eV on the Pt3Co–N12 surface, which is 0.10 eV lower than that on the Pt3Co–N12 surface, indicating the higher reactivity of the Pt3Co–N12 surface. Bader charge analysis in combination with the charge density difference revealed the electron transference of such key elementary step on both two surfaces. For the CO* intermediate on the Pt3Co–N8 surface, the electrons transfer from the Pt3Co cluster to the CO* fragment with the Bader charge value (Δq) calculated to be 0.20 |e|. With the CO* electrocatalytic oxidation to COOH*, electrons transferred from the Pt3Co cluster to the COOH* fragment with valence electron accumulation of 0.35 |e|, whereas on the Pt3Co–N12 surface, electron transferences from the substrate to the two adsorbates were calculated to be 0.11 |e| and 0.19 |e| for the two adsorbates, respectively, which is less than those on the Pt3Co–N8 surface. These results indicate that the higher N coordination number leads to a higher electron share from the Pt3Co cluster, therefore facilitating electron accumulation on the active sites and enhancing the MOR reactivity.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4sc07905a |
This journal is © The Royal Society of Chemistry 2025 |