Wenxin
Guo‡
ab,
Xiaoping
Gao‡
a,
Mengzhao
Zhu‡
a,
Chenxi
Xu
c,
Xiaorong
Zhu
d,
Xuyan
Zhao
a,
Rongbo
Sun
a,
Zhenggang
Xue
e,
Jia
Song
a,
Lin
Tian
a,
Jie
Xu
f,
Wenxing
Chen
g,
Yue
Lin
h,
Yafei
Li
i,
Huang
Zhou
*a and
Yuen
Wu
*ab
aDepartment of Chemistry, iChEM (Collaborative Innovation Center of Chemistry for Energy Materials), University of Science and Technology of China, Hefei 230026, China. E-mail: yuenwu@ustc.edu.cn
bDalian National Laboratory for Clean Energy, Dalian 116023, China
cSchool of Materials Science and Engineering, Hefei University of Technology, Hefei 230009, China
dSchool of Chemistry and Chemical Engineering, Nantong University, Nantong 226019, China
eNEST Laboratory, Department of Physics, Department of Chemistry, College of Science, Shanghai University, Shanghai 200444, China
fInstitute of Functional Nano and Soft Materials (FUNSOM), Jiangsu Key Laboratory for Carbon-Based Functional Materials & Devices, Soochow University, Suzhou 215123, China
gBeijing Key Laboratory of Construction Tailorable Advanced Functional Materials and Green Applications, School of Materials Science and Engineering, Beijing Institute of Technology, Beijing 100081, China
hHefei National Research Center for Physical Sciences at the Microscale, University of Science and Technology of China, Hefei 230026, China
iJiangsu Collaborative Innovation Centre of Biomedical Functional Materials, School of Chemistry and Materials Science, Nanjing Normal University, Nanjing 210023, China
First published on 28th November 2022
Diminishing the usage of Pt without sacrificing its activity still remains a challenge in proton-exchange membrane fuel cells (PEMFCs). Here, we report a gas-promoted dealloying process to prepare a closely packed hybrid electrocatalyst containing Pt-based alloy nanocrystals (NCs) and dense isolated Ni sites. Driven by ammonia and heat, the initial Pt1.5Ni NC undergoes a dealloying process to form a stable Pt-skin Pt1.5Ni1−x alloy due to the continuous detachment of Ni atoms from it. Subsequently, these Ni atoms would be trapped by the adjacent defects on the carbon substrates, resulting in abundant Ni sites distributed closely around the dealloyed Pt1.5Ni1−x NC. For a multielectron transferred oxygen reduction reaction (ORR), the hybrid ensures the reduction of the two electrons at Ni single sites, and the corresponding intermediate (OOH*) rapidly migrates to the neighboring Pt-based NC to finish the subsequent electron transfer. This efficient relay catalytic process could greatly reduce the usage of Pt. The resulting catalyst exhibits excellent ORR activity with a mass activity (MA) of 4.10 A mgPt−1, exceeding that of commercial Pt/C by a factor of ∼15. More importantly, in practical H2/O2 fuel cell tests, a peak power density of 1.72 W cm−2 and a current density of 0.55 A cm−2 at 0.80 V can be achieved, both of which exceed DOE 2025 targets.
Broader contextAs a promising clean and efficient energy carrier, hydrogen is one of the best options to achieve energy technology revolution and large-scale decarbonization. Fuel cell electric vehicles (FCVs), a common terminal application of hydrogen energy, are expected to be an important electric transportation solution by converting the chemical energy of hydrogen into electrical energy through proton exchange membrane fuel cells (PEMFCs). The U.S. Department of Energy has set long-term goals for heavy-duty FCVs with fuel cell system lifetimes and costs of 30000 hours and $60 kW−1, respectively, which urgently requires the development of low-Pt loaded and highly performance catalysts to catalyze the kinetically sluggish cathodic oxygen reduction reaction (ORR), thereby reducing costs and further improving the economics of FCVs. Here, we integrated PtNi nanocrystals and dense isolated Ni sites on N–C to construct a closely packed hybrid for efficient relay catalysis, which greatly improved membrane electrode assembly (MEA) activity and stability while maintaining a low Pt loading. The rational design of high-efficiency relay catalysts not only offers the possibility of further promoting the development of FCV as automotive power sources, but also provides ideas to explore more electrocatalytic processes. |
To better use M–N–C in the 4e− ORR to reduce the usage of Pt, an efficient approach is to replace only 1-2e− of the reduction process rather than the whole reduction process. Meanwhile, another site is also required to complete the transfer of other electrons by relay catalysis. In such a catalytic process, the intermediate species generated on M–N–C need to migrate promptly to the adjacent site, such as a NC or a cluster, while avoiding direct desorption without engaging in the subsequent reaction. This requires an effective coupling of the two active sites to shorten the diffusion distance of the intermediates during the ORR and carry out a cascade of multistep reactions. The complex sites of adjacent M–N–C and NC/cluster might ensure the complete and efficient reduction of continuously migrating intermediates in the 4e− process and might help to further reduce the effects of mass transfer and concentration polarization during the kinetic process.16,17 Thus, the full utilization of M–N–C by relay catalysis is a promising method to reduce the dosage of Pt as well as the catalyst cost.
Herein, we develop a gas-promoted dealloying process to directly construct a closely packed Pt–Ni alloy NC surrounded by single Ni sites, in which an efficient electron transfer relay can be achieved between these two types of sites to realize superior ORR performance. The obtained Pt–Ni NC possesses a stable platinum-rich surface due to the dealloying effect caused by heat and NH3 corrosion, in which abundant Ni atoms migrated from the initial Pt–Ni alloy and were trapped by adjacent nitrogen defect sites. This evolution process was well traced by aberration corrected high-angle annular dark-field scanning transmission electron microscopy (AC HAADF-STEM) and X-ray absorption fine structure (XAFS) spectroscopy. Compared with the untreated sample, the closely packed hybrid displays an enhanced ORR performance in acidic media with a mass activity (MA)/specific activity (SA) of 4.1 A mgPt−1/4.6 mA cm−2, which is 15/7 times that of commercial Pt/C and exceeds that of many noble metal catalysts. More importantly, the catalyst with only half of the Pt loading of Pt/C exhibits a much higher activity and durability in practical fuel cell tests. Finally, theoretical results reveal that the closely packed hybrid can ensure that the key intermediate species OOH* continuously migrates from the Ni single sites to the neighboring Pt-based NC, contributing to a relay reaction to enhance the catalytic performance.
To investigate the intrinsic evolution mechanism from Pt1.5Ni/N–C to Pt1.5Ni1−x/Ni–N–C, representative atomic resolution HAADF-STEM images were acquired under NH3 with 50 °C as the reaction temperature interval. Fig. 2(a) reflects that the synthesized Pt1.5Ni NPs possess an octahedral morphology at room temperature with a lattice stripe of 0.23 nm, consistent with the (111) plane (Fig. S8, ESI†). The elemental mappings further reveal that the Pt1.5Ni NPs possess Pt-rich axes and corners and Ni-rich facet compositions (Fig. 2(a) and Fig. S9, ESI†). This can also be demonstrated by the line scan oriented along the 〈100〉 zone axes, which confirms the start signal with Ni enrichment in facets and Pt enrichment in central axes (Fig. 2(b)). At 50 °C, the initial octahedral morphology is almost unchanged, reflecting good structural stability (Fig. 2(c)). As the temperature increased to 100 °C and 200 °C, the particles gradually eroded into a concave shape, indicating that a dealloying process occurred on the surface. This phenomenon is due to the effect of NH3 molecules, resulting in the leaching of Ni atoms from the alloy particles and their eventual capture by N defects (Fig. 2(d) and (e)). Meanwhile, abundant Pt atoms migrated from Pt-rich edges to the vacant (111) facets under the sustained chaotic thermal motion of molecules. As the temperature rose further, the dealloying process intensified, resulting in the transformation of a truncated octahedron shape with a newly formed (100) plane, accompanied by abundant Ni single sites on the adjacent N–C (Fig. 2(f) and Fig. S10–S12, ESI†). At 300 °C, the truncated octahedron was further converted into a spherical-like shape, which was unchanged even at 350 °C, revealing good thermal stability (Fig. 2(g) and Fig. S13, ESI†).22,23 Density functional theory (DFT) calculations were performed to further probe the evolution mechanism. The results show the migration of a single Ni atom by NH3 coordination, and the subsequent capture by N4 defects is exothermic and requires a diffusion potential barrier of 2.85 eV to reach stable states, which is much smaller than the counterpart of Pt (3.78 eV, Fig. 2(h) and Fig. S14, ESI†). This further demonstrates that Ni atoms migrate more easily than Pt atoms in the bulk phase, agreeing well with the previous STEM-EDX observations (Fig. 1(f)–(h)).
Powder X-ray diffraction (XRD) patterns were adopted to investigate the physical phase of Pt1.5Ni/N–C before and after treatment (Fig. 3(a)). The peak position corresponding to the (111) plane of the face-centered cubic (fcc) Pt–Ni alloy shifts slightly towards the lower 2θ degree of the Pt position, which could be attributed to the lattice spacing expansion caused by Ni leaching from the bulk phase of Pt1.5Ni.24 For Ni X-ray photoelectron spectroscopy (XPS) spectra (Fig. 3(b)), the relative peak ratio of Nix+(0 < x < 3) 2p3/2 in Pt1.5Ni1−x/Ni–N–C is significantly increased compared to Pt1.5Ni/N–C, suggesting that the partial Ni was oxidized in Pt1.5Ni1−x/Ni–N–C, which can also be confirmed by the near edge X-ray absorption fine structure (NEXAFS) spectra of Ni L-edge (Fig. S15, ESI†). The Pt1.5Ni1−x/Ni–N–C shows a weakened intensity due to the separation of the metallic phase. Alternatively, the collected Pt 4f XPS spectra in both Pt1.5Ni/N–C and Pt1.5Ni1−x/Ni–N–C indicate that the surface Pt is mainly in the metallic state (Fig. S16, ESI†).25 The high-resolution N 1s XPS spectra reveal that both Pt1.5Ni1−x/Ni–N–C and Pt1.5Ni/N–C possess pyridyl-N (398.5 eV), pyrrole-N (400.5 eV), metal-N (399.4 eV) and graphitic-N (401.4 eV),26 while the enhanced peak is detected at 399.4 eV in the former due to the generation of the Ni–N species (Fig. S17, ESI†). The N K-edge NEXAFS further shows the existence of π* excitation of pyridyl-N (398.9 eV) and graphitic-N (401.9 eV) species and σ* excitation of the C–N moieties (406.7 eV),27 illustrating that the samples were well graphitized (Fig. S18, ESI†). Moreover, the Raman spectra show an increased ID/IG value (the intensity ratio of the D-band and G-band, 0.93 vs. 0.95) after treatment, demonstrating the generation of carbon defects (Fig. S19, ESI†).
NEXAFS and extended X-ray absorption fine structure (EXAFS) spectroscopy was performed to trace the changes in electronic states and coordination environment during the treatment. With the annealing temperature increasing, the Ni K-edge NEXAFS (Fig. 3(c)) spectra show that the white line (WL) intensity gradually increased and the absorption threshold position shifted toward Ni3+, indicating that Ni was further oxidized. Moreover, the EXAFS spectra of the R space show that the Ni–Ni bond had a certain degree of positive shift with respect to the Ni foil, and a new peak gradually emerged at ∼1.5 Å with lower intensity (Fig. 3(d)), revealing the gradual decrease in surface metallicity and the formation of Ni single sites. The fitting results of Pt1.5Ni1−x/Ni–N–C reveal that the peak located at 1.49 Å can be assigned to Ni–N bonds, and the corresponding coordinated number is 4. For Pt L3-edge XANES spectra, WL strength was enhanced gradually during annealing, which is related to the electron transfer within the alloy, further evidencing the decrease in alloying (Fig. 3(e)). More detailed fitting curves and parameters are displayed in the ESI,† Table S1.
To verify the catalytic performance of the hybrid (Pt1.5Ni1−x/Ni–N–C) obtained by the gas-promoted dealloying process, an ORR experiment was performed. The linear sweep voltammetry (LSV) curve indicates that the half-wave potential (E1/2) is 0.967 V for Pt1.5Ni1−x/Ni–N–C in 0.1 M HClO4 solution, which is superior to the untreated Pt1.5Ni/N–C (0.926 V), the commercial Pt/C (0.887 V) and most of the reported Pt-based catalysts (Fig. 4(a) and Table S2, ESI†). However, the single Ni site catalyst (Ni–N–C) prepared via an impregnation method (ESI†) shows a much poorer activity (E1/2 = 0.38 V). The fast ORR reaction kinetics and excellent activity of Pt1.5Ni1−x/Ni–N–C can be further observed by the small Tafel slope value Pt1.5Ni1−x/Ni–N–C (∼55.44 mV dec−1) < Pt1.5Ni/N–C (∼64.83 mV dec−1) < Pt/C (∼70.46 mV dec−1) < Ni–N–C (∼132.50 mV dec−1) (Fig. S20, ESI†).28 These results illustrate that the design of closely packed hybrids to obtain adjacent active sites is critical for acquiring a high ORR performance. The cyclic voltammetry (CV) curves were measured in a N2-saturated electrolyte (Fig. 4(b)) to obtain the electrochemically active surface area (ECSA) value of Pt1.5Ni1−x/Ni–N–C, which was estimated to be 89.88 m2 gPt−1, greater than that of Pt1.5Ni/N–C (76.07 m2 gPt−1), revealing that the former has a higher utilization of platinum.17,29 Combined with the almost unchanged surface area before and after treatment (766.9 m2 g−1 for Pt1.5Ni/N–C and 785.828 m2 g−1 for Pt1.5Ni1−x/Ni–N–C), it is reasonable to infer that Pt1.5Ni1−x/Ni–N–C possesses an increased number of Pt active sites (Fig. S21, ESI†). Furthermore, the rotating ring-disk electrode (RRDE) test shows an electron transfer number of ∼3.95 and a hydrogen peroxide yield of ∼4% for Pt1.5Ni1−x/Ni–N–C, suggesting a four-electron mechanism similar to that of commercial Pt/C (Fig. S22, ESI†). However, Ni–N–C displays an over 15% hydrogen peroxide yield with a much lower electron transfer number (∼3.4). This result suggests that the closely packed hybrid of Pt1.5Ni1−x/Ni–N–C can effectively inhibit the low electron transfer reaction on the Ni–N–C support, resulting in a superior catalytic performance. Additionally, Pt1.5Ni1−x/Ni–N–C realizes the optimal mass activities (MA) and specific activities (SA) under different potentials, especially for 4.1 AmgPt−1/4.6 mA cm−2 at 0.9 V, which exceeds that of Pt1.5Ni/N–C by a factor of ∼2.9/2.5 and Pt/C by a factor of ∼15.2/7.0 (Fig. 4(c) and (d)). The accelerated deterioration test (ADT) demonstrates that Pt1.5Ni1−x/Ni–N–C retains an outstanding stability after cycling with no observable decay of E1/2. The MA and SA of Pt1.5Ni1−x/Ni–N–C showed no obvious attenuation after 10000 cycles, while the catalyst showed a certain degree of attenuation at high potential and 20000 cycles. (Fig. 4(e) and Fig. S23–S25, ESI†). TEM images of Pt1.5Ni1−x/Ni–N–C after ADT manifest negligible changes in particle morphology and size (Fig. S26, ESI†), further indicating the good structural stability of the catalyst. In contrast, Pt/C shows a severe decrease in E1/2 after only 5000 cycles (Fig. S27 and S28, ESI†).
Moreover, to testify the practical performance of the prepared catalyst Pt1.5Ni1−x/Ni–N–C, we incorporated it into a membrane electrode assembly (MEA) as the cathode catalyst with a Pt loading of 0.05 mgPt cm−2 and assembled it into a PEMFC single cell. The current–voltage (i–V) polarization curves and power density distribution plots of Pt1.5Ni1−x/Ni–N–C demonstrate maximum power densities of 1.23, 1.42, 1.52 and 1.72 Wcm−2 in the H2/O2 cell under partial pressures of 0.0, 0.5, 1.0 and 2.0 bar, respectively (Fig. 4(f)), superior to Pt1.5Ni/N–C and commercial 20 wt% Pt/C MEA tested under the same conditions (Fig. S29 and S30, ESI†). Meanwhile, the through entire polarization scanning current densities of Pt1.5Ni1−x/Ni–N–CNi1−x/Ni–N–C MEA are higher than those of Pt1.5Ni/N–C and Pt/C MEA. For example, the current density for Pt1.5Ni1−x/Ni–N–C at the high voltage region (0.8 V) can reach 0.55 Acm−2, outperforming that for Pt1.5Ni/N–C (0.37 Acm−2) and Pt/C (0.34 Acm−2) at 2.0 bar. Pt1.5Ni1−x/Ni–N–C also exhibits maximum power densities of 0.91 and 0.82 Wcm−2 under different partial pressures and higher current densities compared with Pt1.5Ni/N–C and Pt/C across the entire polarization scan in H2–air cell tests (Fig. 4(g) and Fig. S31, S32, ESI†). In addition, Pt1.5Ni1−x/Ni–N–C MEA and Pt/C MEA were subjected to ADT experiments according to the U.S. DOE (Department of Energy) protocols.30 The i–V polarization curves of Pt1.5Ni1−x/Ni–N–C MEA show no significant decay in current density after 30000 continuous cycles and exhibit a low voltage drop of only 12 mV at a current density of 0.8 A cm−2, indicating a better MEA stability than Pt1.5Ni/N–C and Pt/C (18 mV and 130 mV decay at 0.8 A cm−2, respectively) (Fig. 4(h) and Fig. S33, S34, ESI†). Additionally, Fig. 4(i) reveals that Pt1.5Ni1−x/Ni–N–C MEA delivers a higher initial MA of 0.70 AmgPt−1 than Pt1.5Ni/N–C and only 26.3% decay after 30000 cycles, achieving the DOE 2025 target (0.44 AmgPt−1 for MA and <40% attenuation after 30000 cycles) (Fig. 4(i) and Fig. S35, ESI†). After 30000 cycles, AC HAADF-STEM and EDS mapping showed that the Pt1.5Ni1−x particles retained the original morphology and composition of the alloy, and the Ni sites were still isolated and dispersed on the NC substrate around the particles (Fig. S36 and S37, ESI†). To recap, the results gleaned above demonstrate the superior activity of Pt1.5Ni1−x/Ni–N–C in both RDE measurements and single fuel cell tests and perfectly cater to envision well-designed efficient PEMFC catalysts. This may be attributed to the relay transfer of reaction intermediates between Pt–Ni NC and Ni single sites, which can be further elucidated by the next theoretical calculations (Fig. 5).
To figure out the mechanism of the outstanding ORR activity observed on Pt1.5Ni1−x/Ni–N–C, comprehensive density functional theory (DFT) calculations were performed. According to the above physical characterization results and the consideration of the computational cost, the Pt1.5Ni1−x/Ni–N–C model consisting chiefly of a small PtNi cluster and its neighboring Ni–N–C sites was established to explicitly consider their roles in improving the ORR activities. The models of the Pt1.5Ni/N–C and Ni–N–C samples were also constructed for comparison (Fig. S38, ESI†). Based on the calculated free energy profile of the reaction path, the rate determining steps (RDS) of Pt1.5Ni1−x/Ni–N–C, Pt1.5Ni/N–C, and Ni–N–C are *OH + H+ + e− → * + H2O, *O + H+ + e− → *OH, and * + O2 + H+ + e− → *OOH, respectively. Their corresponding largest free energy changes are 0.31, 0.35, and 0.81 eV at 0.9 V, indicating that Pt1.5Ni1−x/Ni–N–C and Pt1.5Ni/N–C are more active in the four-electron ORR path than Ni–N–C. Meanwhile, the PtNi cluster sites show a much lower free energy change than that of the Ni–N–C sites, which is indicative of more inclination to the four-electron ORR path on the Pt1.5Ni1−x/Ni–N–C catalyst. The path shown by the yellow arrow in Fig. 5(b) and the pink line in Fig. 5(a) is path A, which occurs at the Pt NC site (Pt site) in Pt1.5Ni1−x/Ni–N–C, where O2 completes the reduction of 4 electrons directly. Due to the trace H2O2 products that have been proven in the RRDE evaluation (Fig. S22, ESI†), the pathway to produce H2O2 is further taken into consideration. The produced H2O2 does not bind to the distant Ni–N–C site (PGM-free site) and can be released and migrate to the PtNi cluster sites in the vicinity, as marked by the blue arrow in Fig. 5(a) and purple arrow in Fig. 5(b) (path C). Then, the H2O2 will be rapidly reduced to H2O on the PtNi cluster sites, which would effectively alleviate the damage of free radicals to carbon components and proton membranes. Interestingly, when the ORR process occurs on the close Ni–N–C site (neighboring the PtNi cluster site) of Pt1.5Ni1−x/Ni–N–C, the *OOH intermediate cannot be stably bound to the close Ni–N–C site and will migrate to the neighboring PtNi site and decompose into O* and *OH (Fig. 5(b) and Fig. S39, ESI†), as shown by the green line in Fig. 5(a) and red arrow route in Fig. 5(b) (path B). This mutual assistance of the PtNi cluster and Ni–N–C together completes a four-electron transfer path. The multisite synergy between the Ni–N–C sites and the PtNi cluster sites effectively activates oxygen molecules, which is beneficial to the further reduction process. Therefore, the catalytic efficiency of the multisite synergistic relay reaction path of Pt1.5Ni1−x/Ni–N–C is beyond those of Pt1.5Ni/N–C and Ni–N–C.
The density of states (DOS) of Pt1.5Ni1−x/Ni–N–C, Pt1.5Ni/N–C, and Ni–N–C as well as their corresponding d-band centers of the active atoms are subsequently calculated to understand the origin of the enhanced ORR activity of Pt1.5Ni1−x/Ni–N–C (Fig. 5(c) and Fig. S40, ESI†). The DOS in Fig. 5(c) shows that the Pt1.5Ni1−x/Ni–N–C and Pt1.5Ni/N–C catalysts have abundant electronic states near the Fermi-level in comparison with that of Ni–N–C, implying a greater charge transfer ability. Moreover, more abundant electronic states near the Fermi-level over Pt1.5Ni1−x/Ni–N–C than over Pt1.5Ni/N–C indicate the possible electron interaction between the PtNi clusters and the Ni–N–C sites, which helps to achieve an enhanced catalytic performance. In addition, the d band center of Pt and Ni in Pt1.5Ni1−x/Ni–N–C is closer to the Fermi-level than that of Ni in Ni–N–C (Fig. S40, ESI†), indicating the existence of electron interactions between the PtNi clusters and the Ni–N–C sites. This result also means stronger adsorption of intermediates on Pt1.5Ni1−x/Ni–N–C than on Ni–N–C, which is consistent with the calculated free energy profile.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2ee02381d |
‡ These authors contributed equally. |
This journal is © The Royal Society of Chemistry 2023 |