CHEMISTRY Copyright © 2020 N-induced lattice …Cai et al., ci. dv. 2020 6 : eaaw8113 3 January...

9
Cai et al., Sci. Adv. 2020; 6 : eaaw8113 3 January 2020 SCIENCE ADVANCES | RESEARCH ARTICLE 1 of 8 CHEMISTRY N-induced lattice contraction generally boosts the hydrogen evolution catalysis of P-rich metal phosphides Jinyan Cai 1 *, Yao Song 1 *, Yipeng Zang 1 , Shuwen Niu 1 , Yishang Wu 1 , Yufang Xie 1 , Xusheng Zheng 2 , Yun Liu 1 , Yue Lin 1† , Xiaojing Liu 1† , Gongming Wang 1† , Yitai Qian 1 P-rich transition metal phosphides (TMPs) with abundant P sites have been predicted to be more favorable for hydrogen evolution reaction (HER) catalysis. However, the actual activities of P-rich TMPs do not behave as ex- pected, and the underlying essence especially at the atomic level is also ambiguous. Our structural analysis reveals the inferior activity could stem from the reduced overlap of atomic wave functions between metal and P with the increase in P contents, which consequently results in too strong P-H interaction. To this end, we used N-induced lattice contraction to generally boost the HER catalysis of P-rich TMPs including CoP 2 , FeP 2 , NiP 2 , and MoP 2 . Refined structural characterization and theoretical analysis indicate the N-P strong interaction could increase the atomic wave function overlap and eventually modulate the H adsorption strength. The concept of lattice engineering offers a new vision for tuning the catalytic activities of P-rich TMPs and beyond. INTRODUCTION Water electrolysis, preferably powered by sustainable energy sources, has been considered as one of the most feasible approaches to achieve scalable and renewable hydrogen production (14). However, the sluggish kinetics and the associated substantial energy losses severely impede the energy conversion efficiency, which imperatively demand efficient water electrolysis catalysts (57). Pt-based materials, as the state-of-the-art hydrogen evolution reaction (HER) catalysts, are seriously constrained by the crustal scarcity and high cost (89). To this end, many researchers have been dedicated to exploring in- expensive and earth-abundant non-noble metal-based HER catalysts, such as transition metal oxides (1012), phosphides (1316), chalco- genides (1719), nitrides (2022), and carbides (2325). Although notable processes have been achieved in the past decades, their performance is still far less than that of Pt, especially at acidic con- dition, and searching new catalysts is still the primary goal (26). Very recently, transition metal phosphides (TMPs) attract special attention for HER catalysis, due to their diverse compositions, unique catalytic surface, excellent chemical stability, and electrical conductivity (127). Typically, the negatively charged P sites with moderate electronegativity could act as Lewis bases to trap the posi- tively charged protons, potentially endowing the TMPs with a Gibbs free energy of adsorbed H (H ad ) close to zero. For this reason, it has been consciously predicted that the HER performance of TMPs could be further boosted with the increase in P contents in the TMPs (2830). For instance, Pan et al. (31) found that CoP nanoparticles exhibited significantly higher catalytic activity than Co 2 P. However, as ratio of P was further increased, the HER performance of TMPs was not promoted as expected (132). The reason for the inferior HER activities of the P-rich phosphides is actually still ambiguous, espe- cially at the atomic level. Undoubtedly, unraveling the underlying catalytic behaviors of P-rich TMPs is the presupposition for essen- tially unlocking the limiting factors for HER catalysis and eventually constructing superior HER catalysts, but unfortunately they have been rarely studied so far. Here, we take CoP 2 , a typical P-rich metal phosphide, as an example to investigate the catalytic performance of P-rich TMPs. According to the bond order conservation principle, the increase in P contents will correspondingly reduce the overlap of atomic wave functions, which typically results in strong interaction between P-rich TMPs and H ad and eventually hinders the intrinsic HER activities. So, the limitation of P-rich TMPs for HER catalysis is probably attributed to the decrease in intrinsic per-site activity. Structural analysis reveals that CoP 2 has a unique crystal structure, where each Co is coordinated with six P atoms to form a slightly distorted octahedron (Fig. 1A), and the distorted octahedrons are connected by the PP covalent bond illustrated by the electron density map in Fig. 1B. Therefore, from the structural perspective, modulating the nonpolar P-P inter- action to tailor the overlap of atomic wave functions is the key to further boost the intrinsic HER activity of CoP 2 . In this work, we rationally incorporate nitrogen into the CoP 2 structures to increase the overlap of atomic wave functions via the N-induced lattice contraction. Excitingly, the N-modified CoP 2 (N-CoP 2 ) nanowires (NWs) display an impressive overpotential of 38 mV at 10 mA cm −2 , which is substantially smaller than the CoP 2 NWs and even close to the benchmark Pt/C catalysts. Refined structural characterization and theoretical calculations consistently reveal that the created PN bonds in N-CoP 2 could bring lattice contraction to increase the atomic wave function overlap, which consequently broadens the local energy bandwidth with the downshift of the valence band center and correspondingly decreases the electron coupling be- tween CoP 2 and H ad to facilitate the HER catalysis. We also demon- strate that the N-P strong interaction can generally boost the HER activities of a variety of P-rich TMPs including CoP 2 , FeP 2 , NiP 2 , and MoP 2 . Using lattice engineering for band structure tuning pro- vides a new pathway to the rational construction of superior HER catalysts and beyond. The N-CoP 2 NWs are fabricated via a three-step process, as illustrated in Fig. 2A. Typically, Co (CO 3 ) x OH y NWs are first grown 1 Hefei National Laboratory for Physical Science at the Microscale, Department of Applied Chemistry, University of Science and Technology of China, Hefei, Anhui 230026, P. R. China. 2 National Synchrotron Radiation Laboratory, University of Science and Technology of China, Hefei, Anhui 230029, P. R. China. *These authors contributed equally to this work. †Corresponding author. Email: [email protected] (G.W.); [email protected] (X.L.); [email protected] (Y.Lin) Copyright © 2020 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original U.S. Government Works. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC). on August 17, 2020 http://advances.sciencemag.org/ Downloaded from

Transcript of CHEMISTRY Copyright © 2020 N-induced lattice …Cai et al., ci. dv. 2020 6 : eaaw8113 3 January...

Page 1: CHEMISTRY Copyright © 2020 N-induced lattice …Cai et al., ci. dv. 2020 6 : eaaw8113 3 January 2020SCIENCE ADANCES | RESEARCH ARTICLE 1 of 8 CHEMISTRY N-induced lattice contraction

Cai et al., Sci. Adv. 2020; 6 : eaaw8113 3 January 2020

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

1 of 8

C H E M I S T R Y

N-induced lattice contraction generally boosts the hydrogen evolution catalysis of P-rich metal phosphidesJinyan Cai1*, Yao Song1*, Yipeng Zang1, Shuwen Niu1, Yishang Wu1, Yufang Xie1, Xusheng Zheng2, Yun Liu1, Yue Lin1†, Xiaojing Liu1†, Gongming Wang1†, Yitai Qian1

P-rich transition metal phosphides (TMPs) with abundant P sites have been predicted to be more favorable for hydrogen evolution reaction (HER) catalysis. However, the actual activities of P-rich TMPs do not behave as ex-pected, and the underlying essence especially at the atomic level is also ambiguous. Our structural analysis reveals the inferior activity could stem from the reduced overlap of atomic wave functions between metal and P with the increase in P contents, which consequently results in too strong P-H interaction. To this end, we used N-induced lattice contraction to generally boost the HER catalysis of P-rich TMPs including CoP2, FeP2, NiP2, and MoP2. Refined structural characterization and theoretical analysis indicate the N-P strong interaction could increase the atomic wave function overlap and eventually modulate the H adsorption strength. The concept of lattice engineering offers a new vision for tuning the catalytic activities of P-rich TMPs and beyond.

INTRODUCTIONWater electrolysis, preferably powered by sustainable energy sources, has been considered as one of the most feasible approaches to achieve scalable and renewable hydrogen production (1–4). However, the sluggish kinetics and the associated substantial energy losses severely impede the energy conversion efficiency, which imperatively demand efficient water electrolysis catalysts (5–7). Pt-based materials, as the state-of-the-art hydrogen evolution reaction (HER) catalysts, are seriously constrained by the crustal scarcity and high cost (8, 9). To this end, many researchers have been dedicated to exploring in-expensive and earth-abundant non-noble metal-based HER catalysts, such as transition metal oxides (10–12), phosphides (13–16), chalco-genides (17–19), nitrides (20–22), and carbides (23–25). Although notable processes have been achieved in the past decades, their performance is still far less than that of Pt, especially at acidic con-dition, and searching new catalysts is still the primary goal (26). Very recently, transition metal phosphides (TMPs) attract special attention for HER catalysis, due to their diverse compositions, unique catalytic surface, excellent chemical stability, and electrical conductivity (1, 27). Typically, the negatively charged P sites with moderate electronegativity could act as Lewis bases to trap the posi-tively charged protons, potentially endowing the TMPs with a Gibbs free energy of adsorbed H (Had) close to zero. For this reason, it has been consciously predicted that the HER performance of TMPs could be further boosted with the increase in P contents in the TMPs (28–30). For instance, Pan et al. (31) found that CoP nanoparticles exhibited significantly higher catalytic activity than Co2P. However, as ratio of P was further increased, the HER performance of TMPs was not promoted as expected (1, 32). The reason for the inferior HER activities of the P-rich phosphides is actually still ambiguous, espe-cially at the atomic level. Undoubtedly, unraveling the underlying

catalytic behaviors of P-rich TMPs is the presupposition for essen-tially unlocking the limiting factors for HER catalysis and eventually constructing superior HER catalysts, but unfortunately they have been rarely studied so far.

Here, we take CoP2, a typical P-rich metal phosphide, as an example to investigate the catalytic performance of P-rich TMPs. According to the bond order conservation principle, the increase in P contents will correspondingly reduce the overlap of atomic wave functions, which typically results in strong interaction between P-rich TMPs and Had and eventually hinders the intrinsic HER activities. So, the limitation of P-rich TMPs for HER catalysis is probably attributed to the decrease in intrinsic per-site activity. Structural analysis reveals that CoP2 has a unique crystal structure, where each Co is coordinated with six P atoms to form a slightly distorted octahedron (Fig. 1A), and the distorted octahedrons are connected by the P─P covalent bond illustrated by the electron density map in Fig. 1B. Therefore, from the structural perspective, modulating the nonpolar P-P inter-action to tailor the overlap of atomic wave functions is the key to further boost the intrinsic HER activity of CoP2.

In this work, we rationally incorporate nitrogen into the CoP2 structures to increase the overlap of atomic wave functions via the N-induced lattice contraction. Excitingly, the N-modified CoP2 (N-CoP2) nanowires (NWs) display an impressive overpotential of 38 mV at 10 mA cm−2, which is substantially smaller than the CoP2 NWs and even close to the benchmark Pt/C catalysts. Refined structural characterization and theoretical calculations consistently reveal that the created P─N bonds in N-CoP2 could bring lattice contraction to increase the atomic wave function overlap, which consequently broadens the local energy bandwidth with the downshift of the valence band center and correspondingly decreases the electron coupling be-tween CoP2 and Had to facilitate the HER catalysis. We also demon-strate that the N-P strong interaction can generally boost the HER activities of a variety of P-rich TMPs including CoP2, FeP2, NiP2, and MoP2. Using lattice engineering for band structure tuning pro-vides a new pathway to the rational construction of superior HER catalysts and beyond.

The N-CoP2 NWs are fabricated via a three-step process, as illustrated in Fig. 2A. Typically, Co (CO3)xOHy NWs are first grown

1Hefei National Laboratory for Physical Science at the Microscale, Department of Applied Chemistry, University of Science and Technology of China, Hefei, Anhui 230026, P. R. China. 2National Synchrotron Radiation Laboratory, University of Science and Technology of China, Hefei, Anhui 230029, P. R. China.*These authors contributed equally to this work.†Corresponding author. Email: [email protected] (G.W.); [email protected] (X.L.); [email protected] (Y.Lin)

Copyright © 2020 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original U.S. Government Works. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC).

on August 17, 2020

http://advances.sciencemag.org/

Dow

nloaded from

Page 2: CHEMISTRY Copyright © 2020 N-induced lattice …Cai et al., ci. dv. 2020 6 : eaaw8113 3 January 2020SCIENCE ADANCES | RESEARCH ARTICLE 1 of 8 CHEMISTRY N-induced lattice contraction

Cai et al., Sci. Adv. 2020; 6 : eaaw8113 3 January 2020

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

2 of 8

on carbon cloth (CC) using a well-developed hydrothermal reaction (33). Then, the prepared Co(CO3)xOHy NWs are converted to CoP2 NWs through a thermal phosphidation process in a home-built tube furnace system. Last, nitrogen is incorporated by calcining the CoP2 NWs in ammonia atmosphere. Figure S1 presents the SEM images of the prepared Co(CO3)xOHy, which show that the entire CC is uniformly covered with well-aligned NW arrays. Even after the phosphidation and the consequent nitrogen incorporation, the NW morphology still can be preserved (Fig. 2B and fig. S2). Meanwhile, the transmission electron microscopy images (fig. S3) indicate that the smooth NW becomes rougher after these thermal treatments. In addition, x-ray diffraction (XRD) is used to acquire the structural information. The diffraction peaks located at around 24.8°, 35.2°, 35.6°, and 37.6° correspond to the (−111), (002), (200), and (−121) planes, respectively, of the monoclinic CoP2 (Joint Committee on Powder Diffraction Standards no. 77-0263). CoP2 and N-CoP2 NWs prepared at 350°C show very similar diffraction patterns, while the diffraction peaks have slightly shifted to the higher angle region with the N in-corporation (Fig. 2C and fig. S4), suggesting that N doping does not

intrinsically change the structures of CoP2 but leads to the slightly compressed interplanar distances. Moreover, with the further in-crease in the ammonia treatment temperatures, N-CoP2 eventually undergoes phase segregation to form the mixture of CoP2 and CoP (fig. S5). The energy-dispersive x-ray (EDX) spectrum of N-CoP2 reveals the atomic ratio of Co, P, and N is around 1:1.74:0.3 (fig. S6). In addition, the high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) and the corresponding EDX elemental mapping images indicate that Co, P, and N are homoge-neously distributed over the NW (Fig. 2D). Meanwhile, the atomic resolution HAADF analysis in Fig. 2 (E and F) reveals that the in-corporation of N into CoP2 results in the decreased d-spacing of the (002) and (020) facets, which is consistent with the peak shift toward higher angles in the XRD patterns. The quantified structure con-traction parameters estimated by atomic resolved HAADF and XRD characterizations are listed in table S1. Moreover, the strain tensor values using the geometric-phase analysis of N-CoP2 are more neg-ative than those of CoP2, proving that N doping gives rise to the lattice contraction in N-CoP2.

Fig. 1. Structure analysis. (A) Crystal structure of CoP2. (B) Electron density distribution between the octahedral units.

Fig. 2. Material synthesis and structural characterization. (A) Schematic illustration of the preparation of N-CoP2 NWs. (B) SEM image of N-CoP2 NWs on CC. (C) XRD patterns of CoP2 and N-CoP2 NWs and the corresponding magnification of the diffraction peaks of the (200) facet. (D) HAADF-STEM image and EDX elemental mapping of Co, P, and N for a single N-CoP2 NW. (E and F) Atomic resolution HAADF images and the corresponding strain tensor mapping and the line profile analysis for CoP2 and N-CoP2, respectively.

on August 17, 2020

http://advances.sciencemag.org/

Dow

nloaded from

Page 3: CHEMISTRY Copyright © 2020 N-induced lattice …Cai et al., ci. dv. 2020 6 : eaaw8113 3 January 2020SCIENCE ADANCES | RESEARCH ARTICLE 1 of 8 CHEMISTRY N-induced lattice contraction

Cai et al., Sci. Adv. 2020; 6 : eaaw8113 3 January 2020

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

3 of 8

To unravel the effects of N dopants on the electronic structures of CoP2, x-ray photoelectron spectroscopy (XPS) is further conducted. Figure 3A presents the typical XPS Co 2p spectra of CoP2 and N-CoP2. The peaks located at ~778.59 and 793.60 eV are ascribed to the Co 2p3/2 and Co 2p1/2 of the Co─P bonds, respectively (32, 34). After N doping at 350°C, the binding energies of Co 2p3/2 and Co 2p1/2 only exhibit a very slight shift to the lower energy region. Moreover, with the increase in the ammonia treatment temperatures, gradual shifts are observed (fig. S7A), which may be originated from the created intrinsic defects, due to the reducing property of the ammonia. Meanwhile, the XPS P 2p spectra in Fig. 3B can be deconvoluted into a pair of peaks, which correspond to the P 2p3/2 and P 2p1/2 of the Co─P bonds (30, 35). Impressively, the peak positions of the P 2p were gradually shifted to the higher binding energy region, with the incorporation of nitrogen (Fig. 3B and fig. S7B), indicating the decreased average electron density around the P sites in N-CoP2, which is probably induced by the N-P strong interaction. Moreover, the broad XPS N 1s spectrum of N-CoP2 can also be deconvoluted into the chemical states of N-Co and N-P (Fig. 3C and fig. S7C), suggesting that N is successfully doped into the CoP2 lattices.

Furthermore, synchrotron-based x-ray absorption fine structure (XAFS) spectroscopy is used to probe the localized coordination environments of the Co sites in the N-CoP2. The x-ray absorption near-edge spectra (XANES) of CoP2 and N-CoP2 with standard Co foil and Co2O3 as references are shown in fig. S8A. The pre-edge of CoP2 slightly shifts to the lower energy values after N doping, which suggests the increased average electron densities around the Co sites. In addition, Co L-edge XANES for surface chemical electronic state characterization is also measured using soft x-ray (fig. S8B), in which the reduced intensity of the L3-edge peak of N-CoP2 also suggests more electron filling of the Co d orbital (36). This is in agreement with the XPS Co 2p and XANES Co K-edge results. The similar oscillation profiles of the extended x-ray absorption fine structure (EXAFS) analysis of k2-weighted (k) signals (fig. S8C) indicate that the overall crystal structure of CoP2 is well maintained after N doping,

consistent with the XRD results in Fig. 2C. Figure 3D shows the corresponding R space curves of Co after k2[(k)]-weighted Fourier transform, where the single strong shell at around 1.78 Å originated from the Co-P vector (fig. S9A). For N-CoP2, the shell becomes slightly broader, which may be attributed to the formation of a Co-N vector with very similar peak position at around 1.78 Å (fig. S9B) (37). Because of the slight bond length difference between Co-N and Co-P, more-refined first shell analysis of CoP2 and N-CoP2 is further carried out. Figure 3 (E and F) shows the wavelet trans-form (WT) contour plots of the two signals based on Morlet wavelets with optimum resolution at the first shell. Apparently, CoP2 has the maximum intensity at k = 4.95 Å−1, while the maximum of N-CoP2 presents at k = 4.90 Å−1. The small shift of the WT maximum mostly originated from the atomic number difference between P and N (fig. S10) (38). Together, XPS and XAFS analyses further demonstrate that N is substitutionally doped into the structure of CoP2, which can es-sentially manipulate the electronic and coordination structures of CoP2.

The catalytic performance for HER catalysis is evaluated in a typical three-electrode configuration with 0.5 M H2SO4 as the electro-lyte. For the sake of comparison, bare CC, CoP2, and Pt/C are also examined. Figure 4A shows the linear sweep voltammetry (LSV) curves at the scan rate of 5 mV s−1. Impressively, N-CoP2 displays the best electrocatalytic activity with an overpotential of only 38 mV at 10 mA cm−2, which is substantially smaller than those of CC (458 mV) and CoP2 (125 mV), and even close to the performance of the benchmark Pt/C (28 mV). It also represents the best performance toward HER catalysis in acidic conditions among reported Co-based compounds (13, 19, 39). The temperature-dependent activities of N-CoP2 (denoted as N-CoP2-X) with various N concentrations [N/(N + P)] quantified by XPS characterizations are also studied and shown in Fig. 4B, fig. S11, and table S2. The maximum HER activity was achieved at the temperature of 350°C with an N content of 17.1%. With further increase in the temperature and N contents, the HER activity begins to decay, which may be attributed to the unfavorable phase segregation to form CoP (figs. S5 and S12). To

Fig. 3. XPS and XAFS characterization. (A) XPS Co 2p spectra and (B) XPS P 2p spectra of CoP2 and N-CoP2. (C) XPS N 1s spectrum of N-CoP2. (D) k2-weighted FT-EXAFS curves of the Co K-edges for CoP2 and N-CoP2. WT-EXAFS spectra of (E) CoP2 and (F) N-CoP2. a.u., arbitrary units.

on August 17, 2020

http://advances.sciencemag.org/

Dow

nloaded from

Page 4: CHEMISTRY Copyright © 2020 N-induced lattice …Cai et al., ci. dv. 2020 6 : eaaw8113 3 January 2020SCIENCE ADANCES | RESEARCH ARTICLE 1 of 8 CHEMISTRY N-induced lattice contraction

Cai et al., Sci. Adv. 2020; 6 : eaaw8113 3 January 2020

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

4 of 8

further probe the catalytic kinetics, Tafel plots derived from the po-larization curves are provided in Fig. 4C. The measured Tafel slope for the N-CoP2 is 46 mV dec−1, much smaller than the 73 mV dec−1 of the CoP2, which proceeds via the Volmer-Heyrovsky mechanism with Heyrovsky step [Had + H+ + e− → H2] as the rate-determining step. To deeply evaluate the correlation of Tafel slopes and the catalytic kinetics, we investigate the hydrogen oxidization reaction (HOR) mechanism experimentally. The measured Tafel slopes of CoP2 and N-CoP2 for HOR are 207 and 341 mV dec−1, respectively, which proceed with the Heyrovsky-Volmer mechanism and the Heyrovsky step [H2 → Had + H+ + e−] as the rate-determining step (fig. S13 and table S3). Together, all these results consistently demonstrate N-CoP2 has superior HER catalytic kinetics due to the accelerated rate- determining Heyrovsky step with weakened adsorption strength between Had and the catalyst surface (40, 41).

Since the surface area variation may also affect the HER activity, cyclic voltammetry (CV) studies are used to estimate the electro-chemical surface areas by measuring the electrochemical double- layer capacitance (Cdl), as shown in fig. S14C. The calculated Cdl of N-CoP2 is 10.8 mF cm−2, which is larger than the 6.1 mF cm−2 of CoP2. However, the catalytic current of N-CoP2 at 50 mV is 32 times higher than that of CoP2, implying the enhanced HER activity did not mainly stem from the increased surface area. To further probe the intrinsic per-site catalytic activity, turnover frequency (TOF) is plotted in Fig. 4D (42). Apparently, the TOF values of N-CoP2 are consistently larger than those of CoP2 in the whole studied potential region, suggesting that the catalytic surface of CoP2 has been essentially changed after nitrogen incorporation. Furthermore, electrochemical impedance spectroscopy (EIS) is used to explore the interfacial charge transfer kinetics on the catalytic surface. Figure 4E shows the Nyquist plots of CoP2 and N-CoP2 at 100 mV versus Reversible Hydrogen Electrode (RHE), respectively. Both the Nyquist plots present typical semicircle profiles from 100 kHz to 0.01 Hz, which can be well fitted using a simplified Randle circuit model (Fig. 4E,

inset). The fitted charge transfer resistance (Rct) for N-CoP2 (2 ohm) is much smaller than that of CoP2 (13 ohm), suggesting that N dop-ing could accelerate the interfacial charge transfer kinetics. Last, the electrochemical stability of N-CoP2 is also evaluated. Figure 4F shows the chronopotentiometric curve of N-CoP2 at the current density of 20 mA cm−2. The overpotential change (versus RHE) during the chronopotentiometric test goes from −53.2 to −64.6 mV in 30 hours. The decay mainly occurs in the first 12 hours, and then the potential becomes relatively stable. The initial decay could be attributed to the generated bubbles that cannot be efficiently released from the elec-trode surface. Furthermore, the SEM image and XRD pattern are also collected after the stability testing, as shown in the insets of Fig. 4F and fig. S15. Both the NW morphology and the crystal struc-ture can be well maintained after the durability testing, further indi-cating the robustness of N-CoP2 NWs for HER catalysis.

Density functional theory (DFT) is further carried out to study the possible effects of nitrogen doping on the HER activity of CoP2. Figure 5A shows the crystal structures of CoP2 and N-CoP2 with labeled lattice constants. The decreased lattice constants after N incorporation reveal the existence of remarkable lattice contraction, which stemmed from the formation of shorter and stronger N─P bonds as junctions of the octahedral units in the N-CoP2. Meanwhile, the mapping of orbital wave functions in Fig. 5A illustrates the strong N-P interaction results in substantially increased atomic orbital wave function overlap integrals, especially between N and P by the type overlap with the decreased principle quantum number for the involved N atomic orbitals. Moreover, the more localized electron density and the decreased bond lengths in the surface models of N-CoP2 with various facets of (100), (001), and (010) (Fig. 5B and figs. S16 and S17) relative to those of CoP2 further support that the N-P interaction is much stronger than the P-P interaction. The increased atomic wave function overlap integrals spread the large energy difference between the highest and lowest orbital levels in the local band and further broaden the local energy bandwidth (43, 44). Figure 5C illustrates

Fig. 4. Electrochemical HER performance. (A) LSV curves for CC, CoP2, N-CoP2, and Pt/C in 0.5 M H2SO4. (B) The overpotentials at 10 mA cm−2 for CoP2, N-CoP2-250, N-CoP2-350, N-CoP2-450, and N-CoP2-550. (C) The corresponding Tafel plots for CC, CoP2, N-CoP2, and Pt/C. (D) TOF plots of CoP2 and N-CoP2 against the potentials for HER. CPE, constant phase element. (E) Nyquist plots of CoP2 and N-CoP2 at 100 mV versus RHE. (F) Chronopotentiometric curve for N-CoP2 at the current density of 20 mA cm−2. The insets are the SEM images of N-CoP2 before and after the stability test.

on August 17, 2020

http://advances.sciencemag.org/

Dow

nloaded from

Page 5: CHEMISTRY Copyright © 2020 N-induced lattice …Cai et al., ci. dv. 2020 6 : eaaw8113 3 January 2020SCIENCE ADANCES | RESEARCH ARTICLE 1 of 8 CHEMISTRY N-induced lattice contraction

Cai et al., Sci. Adv. 2020; 6 : eaaw8113 3 January 2020

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

5 of 8

how the energy levels and bandwidths vary with the internuclear distance of atoms in a crystalline solid, in which the shaded areas represent the energy bands formed by the valence orbitals. As the internuclear distance decreases, the collection of valence orbitals gradually broadens into bands, where the formed overlapped bands (the intermix) further broaden the bandwidth. As the bandwidth widens, the only way to maintain the number of fixed valence elec-trons is to shift the valence band center of CoP2 away from the Fermi level, which is also reflected by the density of states (DOS) with the shifted valence band center from −1.32 to −1.36 eV (Fig. 5C). Ac-cording to the Newns-Anderson model that expresses the adsorption energies depend on the interactions between the catalyst valence bands and the adsorbate orbitals, the electron coupling between them gives rise to the bonding and antibonding state splitting, as illustrated in Fig. 5C. Since the wave function of high-lying antibonding states has dominant catalyst valence states characters, the downshift of the CoP2 hybridization valence band center will correspondingly reduce the antibonding level relative to the Fermi level (Fig. 5C and fig. S18) and consequently lead to the weak adsorption strength that origi-nated from more electron fillings of antibonding states (fig. S19). We also study the effects of N concentration on the calculated adsorption energies of hydrogen (GH*), the GH* on the various crystal facets of N-CoP2 with different N concentrations are consistently much closer to neutral than those on the CoP2 (Fig. 5D and figs. S20 and S21). Moreover, the interplanar distances of various crystal facets of N-CoP2 are also consistently decreased (table S4). Typically, the GH* close to neutral is the key descriptor for HER catalysis in acidic condition. Therefore, the smaller GH* value induced by the strong N-P inter-action enables N-CoP2 to be highly active for HER catalysis.

Since intrinsic defects may also be created during the ammonia treatment, we further theoretically and experimentally study the effects of the defects on the HER catalysis of CoP2. As shown in fig. S22A, although the GH* value of the defective CoP2 is closer to neutral than that of CoP2, it is still much inferior to that of N-CoP2. Meanwhile, defects in CoP2 are also intentionally introduced by hydrogen treatment and experimentally evaluated for HER catalysis. Electrochemical

characterization indicates that the HER activity of CoP2 can be improved with the introduction of defects (fig. S22B). However, considering that the performance of defective CoP2 (D-CoP2) is still far less than that of N-CoP2, it suggests that the exceptionally high HER activity of the N-CoP2 did not mainly stem from the intrinsic defects but the N-induced lattice contraction.

In view of the existence of P─P bonds in many P-rich TMPs, we also evaluate the feasibility of nitrogen doping for improving the HER catalysis of other P-rich TMPs including of FeP2, NiP2, and MoP2. Figures S23 and S24 show the crystal structures and the pro-jected density of state (PDOS) plots of FeP2, NiP2, MoP2, and their corresponding N-doped counterparts. Similarly, all of them consist-ently undergo lattice contraction with increased atomic wave func-tion overlap integrals, which consequently result in broadened energy bandwidths and the downshift of the valence band centers by the formation of N-P strong interactions. To further prove the feasibility for HER catalysis, we experimentally fabricated these P-rich metal phosphides and studied their electrochemical properties including LSV, Tafel, and stability (figs. S25 to S27). The electrochemical studies consistently indicate that the N doping can substantially improve the HER performances of FeP2, NiP2, and MoP2 with overpotentials of 97, 153, and 135 mV, respectively, at the current of 10 mA cm−2. These results suggest that lattice contraction induced by the N-P strong interactions is a general strategy to boost the HER catalysis of P-rich TMPs.

In summary, we have demonstrated that the HER performance of P-rich CoP2 can be significantly boosted by N-induced lattice contraction. The prepared N-CoP2 exhibits an impressive over-potential of 38 mV at 10 mA cm−2, which is very close to that of precious Pt catalysts in acidic conditions and also represents the best activity among reported Co-based catalysts. Refined structural char-acterization and DFT calculation consistently reveal that the created strong N-P interaction in CoP2 could increase the atomic wave func-tion overlap via the lattice contraction and consequently broaden the local energy bandwidth with the downshift of the valence band center, which eventually facilitate the HER catalysis by weakening the electron

Fig. 5. DFT calculations. (A) Crystal structures of CoP2 and N-CoP2 with lattice constants and the corresponding mapping of orbital wave functions. (B) Surface electron density difference of CoP2 (100) and N-CoP2 (100). (C) DOS of CoP2 and N-CoP2 with/without hydrogen adsorption and the schematic diagram of the relation between the broadened bandwidth and H adsorption strength. (D) Calculated GH* on CoP2 and N-CoP2 with various facets.

on August 17, 2020

http://advances.sciencemag.org/

Dow

nloaded from

Page 6: CHEMISTRY Copyright © 2020 N-induced lattice …Cai et al., ci. dv. 2020 6 : eaaw8113 3 January 2020SCIENCE ADANCES | RESEARCH ARTICLE 1 of 8 CHEMISTRY N-induced lattice contraction

Cai et al., Sci. Adv. 2020; 6 : eaaw8113 3 January 2020

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

6 of 8

coupling between CoP2 and Had. The lattice contraction strategy via N-P strong interactions can generally boost the HER catalysis of a variety of P-rich TMPs including CoP2, FeP2, NiP2, and MoP2, of-fering a new lattice engineering approach to fundamentally tuning the electronic properties of P-rich TMPs for HER catalysis and beyond.

MATERIALS AND METHODSThe synthesis of CoP2 and N-CoP2 NWsCoP2 NWs were fabricated by the thermal phosphidation of Co(CO3)xOHy NWs. The Co(CO3)xOHy NWs were first grown on CC by a conventional hydrothermal reaction (33). Typically, 1 mmol of Co(NO3)2·6H2O, 5 mmol of CO(NH2)2, and 2 mmol NH4F were dissolved in 20 ml of deionized water under vigorous stirring. Then, the prepared homogeneous solution was transferred into a 25-ml Teflon-lined autoclave with a piece of hydrophilic CC placed against the wall of the autoclave. Last, the reaction solution was maintained at 120°C for 4 hours to synthesize the Co(CO3)xOHy NWs on CC. The as-synthesized Co(CO3)xOHy was washed with water and ethanol several times and lastly dried under a vacuum oven. The phosphidation treatment was conducted at 550°C for 3 hours in a tube furnace system with red phosphorus as the P source and argon gas as the carrier gas. N-doped CoP2 NWs (N-CoP2) were achieved by annealing the CoP2 under ammonia condition at various temperatures (250° to 550°C with interval of 100°C) for 20 min.

Synthesis of N-FeP2, N-NiP2, and N-MoP2The P-rich metal phosphides including FeP2, NiP2, and MoP2 were prepared using a similar method used for CoP2. FeOOH, Ni(OH)2, and MoS2 for the following the phosphidation treatment were syn-thesized using previously reported methods (45–47). FeP2 and NiP2 were achieved at the phosphidation temperature of 550°C for 3 hours, while MoP2 was obtained at the temperature of 800°C for 2 hours. N doping for FeP2, NiP2, and MoP2 was performed at 350°C for 30 min, 300°C for 30 min, and 550°C for 20 min, respec-tively, under ammonia condition.

Material characterizationXRD was conducted on a Philips X’Pert Pro Super diffractometer (with Cu K,  = 1.54182 Å). The XPS was performed at the photo-emission end-station (BL10B) in the National Synchrotron Radiation Laboratory, Hefei. Field emission scanning electron microscopy was carried out on the JEOL-JSM-6700F, while the high-resolution trans-mission electron microscopy, HAADF-STEM, and energy-dispersive spectroscopy (EDS) mapping analyses were performed on the JEOL JEM-ARF200F TEM/STEM with a spherical aberration corrector. The XAFS spectra were collected at the 14W1 station in the Shanghai Synchrotron Radiation Facility. The collected EXAFS data were analyzed using the ATHENA program as implemented in the IFEFFIT software packages according to the standard procedures. The k2-weighted EXAFS spectra were achieved by pre-edge background deduction and then normalized relative to the edge-jump step, and the plotting k-weight was 2.

Electrochemical characterizationAll the electrochemical tests were performed on a CHI 760E electro-chemical workstation (CH Instruments, China) at room temperature. The HER catalysis was evaluated in a three-electrode setup with

0.5 M H2SO4 solution as the electrolyte, a standard Ag/AgCl electrode (saturated KCl solution) as the reference electrode, and a graphite rod as the counter electrode. In the electrochemical testing, the measured potentials were calibrated with respect to RHE. The con-version equation is E(RHE) = E(Ag/AgCl) + 0.197 V + 0.059 × PH. The EIS was tested at the overpotential of 100 mV versus RHE, with the frequencies varying from 100 kHz to 0.01 Hz. All potentials were also corrected by eliminating the electrolyte resistances obtained from the EIS spectra. The TOF values were estimated according to the previous reported method (48), and the specific crystal parameters of different materials are listed in table S5. The most active sites are supposed to be P, while the Co and N sites are relatively inactive. Typically, the HER activity is the average results of various sites. Thus, for TOF calculation, we consider all the possible sites as the number of active sites, because such calculation will not overestimate the per-site activity.

Computational detailsDFT calculation was performed using the CASTEP program with ultrasoft pseudopotentials as implemented in the Materials Studios package of Accelrys Inc. (49). The electron exchange-correlation potential was dealt with the Perdew-Burke-Ernzerhof functional of generalized gradient approximation. The core electrons of atoms were treated using effective core potential. The smooth parts of the wave functions were expanded with a kinetic energy cutoff of 400 eV. The 5 × 5 × 5 and 2 × 2 × 1 Monkhorst-Pack mesh k-points were adopted for Brillouin zone sampling of bulk and surface calculations, in which the k-points are 63 and 2, respectively. The convergence criterions were set to 5.0 × 10−6 eV per atom for energy, 5.0 × 10−4 Å for maximum displacement, and 0.01 eV Å−1 for maximum force. The lattice parameters of the CoP2 unit cell after geometry optimization were a = 5.52 Å, b = 5.53 Å, c = 5.59 Å, which were comparable with the experimental data (a = 5.55 Å, b = 5.55 Å, c = 5.61 Å). The surfaces (001), (010), and (100) were modeled by a periodic slab repeated in a 2 by 2 by 2 surface unit cell with a vacuum region of 12 Å between the slabs along the z axis. The whole configuration in all the calcula-tions was relaxed during the geometry optimizations and was carried out in the reciprocal space. The free energy of Had was determined according to the previous report (50). The same calculation parameters as CoP2 were also applied for FeP2, NiP2, and MoP2, and the used crystal parameters of the unit cells are listed in table S6.

SUPPLEMENTARY MATERIALSSupplementary material for this article is available at http://advances.sciencemag.org/cgi/content/full/6/1/eaaw8113/DC1Fig. S1. SEM images of Co(CO3)xOHy NWs collected with different magnifications.Fig. S2. The influence of ammonia treatment temperature on the morphology of N-CoP2-X (X is the temperature).Fig. S3. Structure characterizations of Co(CO3)xOHy, CoP2, and N-CoP2.Fig. S4. The magnified XRD patterns of CoP2 and N-CoP2.Fig. S5. The influence of ammonia treatment temperature on the composition of N-CoP2-X (X is the temperature).Fig. S6. Elemental composition characterization of N-CoP2 NWs.Fig. S7. XPS spectra of Co 2p, P 2p, and N 1s.Fig. S8. XAFS and XANES characterizations of CoP2 and N-CoP2.Fig. S9. The k2-weighted FT-EXAFS curves of CoP2 and N-CoP2 with CoP and CoNx as reference.Fig. S10. WT-EXAFS of CoP and CoNx.Fig. S11. The electrochemical HER performances of N-CoP2-X (X = 250, 350, 450, and 550) in 0.5 M H2SO4.Fig. S12. The influences of N contents on the overpotentials of N-CoP2-X (X = 250, 350, 450, and 550).Fig. S13. The electrochemical HOR performances of CoP2 and N-CoP2 in 0.5 M H2SO4.

on August 17, 2020

http://advances.sciencemag.org/

Dow

nloaded from

Page 7: CHEMISTRY Copyright © 2020 N-induced lattice …Cai et al., ci. dv. 2020 6 : eaaw8113 3 January 2020SCIENCE ADANCES | RESEARCH ARTICLE 1 of 8 CHEMISTRY N-induced lattice contraction

Cai et al., Sci. Adv. 2020; 6 : eaaw8113 3 January 2020

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

7 of 8

Fig. S14. CV curves at different scan rates and the capacitive currents as a function of scan rate in 0.5 M H2SO4.Fig. S15. The XRD patterns of N-CoP2 before and after the durability test.Fig. S16. The electron density differences of CoP2 and N-CoP2.Fig. S17. The optimized structures of CoP2 and N-CoP2 surfaces.Fig. S18. The PDOS of CoP2 and N-CoP2 with different surfaces.Fig. S19. The PDOS of CoP2 and N-CoP2 with the hydrogen adsorbed on different surfaces.Fig. S20. The optimized surface models of CoP2 and N-CoP2 with hydrogen adsorbed on different surfaces.Fig. S21. The calculated GH* on CoP2 and N-CoP2 with various N concentrations on different facets.Fig. S22. The calculated GH* and LSV curves in 0.5 M H2SO4 for CoP2, D-CoP2, and N-CoP2.Fig. S23. The optimized crystal structures of FeP2/N-FeP2, NiP2/N-NiP2, and MoP2/N-MoP2.Fig. S24. The PDOS of FeP2/N-FeP2, NiP2/N-NiP2, and MoP2/N-MoP2.Fig. S25. The morphology characterizations and electrochemical HER performances of FeP2 and N-FeP2.Fig. S26. The morphology characterizations and electrochemical HER performances of NiP2 and N-NiP2.Fig. S27. The morphology characterizations and electrochemical HER performances of MoP2 and N-MoP2.Table S1. HAADF and XRD calculated lattice contraction for (020) and (002) facets.Table S2. N contents acquired from XPS characterization for N-CoP2 with various treatment temperatures.Table S3. Tafel slopes for HOR with various mechanisms.Table S4. The derived strain for (020) and (002) by the DFT calculation with various N contents.Table S5. Crystallographic data of CoP2 and N-CoP2 used for calculating the number of active sites.Table S6. Crystallographic data of FeP2/N-FeP2, NiP2/N-NiP2, and MoP2/N-MoP2.

REFERENCES AND NOTES 1. Y. M. Shi, B. Zhang, Recent advances in transition metal phosphide nanomaterials:

Synthesis and applications in hydrogen evolution reaction. Chem. Soc. Rev. 45, 1529–1541 (2016).

2. X. X. Zou, Y. Zhang, Noble metal-free hydrogen evolution catalysts for water splitting. Chem. Soc. Rev. 44, 5148–5180 (2015).

3. V. R. Stamenkovic, D. Strmcnik, P. P. Lopes, N. M. Markovic, Energy and fuels from electrochemical interfaces. Nat. Mater. 16, 57–69 (2017).

4. I. Roger, M. A. Shipman, M. D. Symes, Earth-abundant catalysts for electrochemical and photoelectrochemical water splitting. Nat. Rev. Chem. 1, 0003 (2017).

5. H. L. Fei, J. C. Dong, Y. X. Feng, C. S. Allen, C. Z. Wan, B. Volosskiy, M. F. Li, Z. P. Zhao, Y. L. Wang, H. T. Sun, P. F. An, W. X. Chen, Z. Y. Guo, C. Lee, D. L. Chen, I. Shakir, M. J. Liu, T. D. Hu, Y. D. Li, A. I. Kirkland, X. F. Duan, Y. Huang, General synthesis and definitive structural identification of MN4C4 single-atom catalysts with tunable electrocatalytic activities. Nat. Catal. 1, 63–72 (2018).

6. M. S. Faber, S. Jin, Earth-abundant inorganic electrocatalysts and their nanostructures for energy conversion applications. Energ. Environ. Sci. 7, 3519–3542 (2014).

7. Y. Jiao, Y. Zheng, M. Jaroniec, S. Z. Qiao, Design of electrocatalysts for oxygen-and hydrogen-involving energy conversion reactions. Chem. Soc. Rev. 44, 2060–2086 (2015).

8. P. T. Wang, X. Zhang, J. Zhang, S. Wan, S. J. Guo, G. Lu, J. L. Yao, X. Q. Huang, Precise tuning in platinum-nickel/nickel sulfide interface nanowires for synergistic hydrogen evolution catalysis. Nat. Commun. 8, 14580 (2017).

9. P. T. Wang, K. Z. Jiang, G. M. Wang, J. L. Yao, X. Q. Huang, Phase and interface engineering of platinum-nickel nanowires for efficient electrochemical hydrogen evolution. Angew. Chem. Int. Ed. 55, 12859–12863 (2016).

10. R. Wu, J. F. Zhang, Y. M. Shi, D. L. Liu, B. Zhang, Metallic WO2-carbon mesoporous nanowires as highly efficient electrocatalysts for hydrogen evolution reaction. J. Am. Chem. Soc. 137, 6983–6986 (2015).

11. Z. W. Fang, L. L. Peng, Y. M. Qian, X. Zhang, Y. J. Xie, J. J. Cha, G. H. Yu, Dual Tuning of Ni-Co-A (A= P, Se, O) Nanosheets by Anion Substitution and Holey Engineering for Efficient Hydrogen Evolution. J. Am. Chem. Soc. 140, 5241–5247 (2018).

12. M. Gong, W. Zhou, M.-C. Tsai, J. G. Zhou, M. Y. Guan, M.-C. Lin, B. Zhang, Y. Hu, D.-Y. Wang, J. Yang, S. J. Pennycook, B.-J. Hwang, H. Dai, Nanoscale nickel oxide/nickel heterostructures for active hydrogen evolution electrocatalysis. Nat. Commun. 5, 4695 (2014).

13. Q. W. Zhou, Z. H. Shen, C. Zhu, J. C. Li, Z. Y. Ding, P. Wang, F. Pan, Z. Y. Zhang, H. X. Ma, S. Y. Wang, H. G. Zhang, Nitrogen-doped CoP electrocatalysts for coupled hydrogen evolution and sulfur generation with low energy consumption. Adv. Mater. 30, 1800140 (2018).

14. J. Q. Tian, Q. Liu, A. M. Asiri, X. P. Sun, Self-supported nanoporous cobalt phosphide nanowire arrays: An efficient 3D hydrogen-evolving cathode over the wide range of pH 0–14. J. Am. Chem. Soc. 136, 7587–7590 (2014).

15. E. J. Popczun, J. R. Mckone, C. G. Read, A. J. Biacchi, A. M. Wiltrout, N. S. Lewis, R. E. Schaak, Nanostructured nickel phosphide as an electrocatalyst for the hydrogen evolution reaction. J. Am. Chem. Soc. 135, 9267–9270 (2013).

16. Y. Li, H. X. Li, K. Z. Cao, T. Jin, X. J. Wang, H. M. Sun, J. X. Ning, Y. J. Wang, L. F. Jiao, Electrospun three dimensional Co/CoP@nitrogen-doped carbon nanofibers network for efficient hydrogen evolution. Energy Storage Mater. 12, 44–53 (2018).

17. T. Y. Kou, T. Smart, B. Yao, I. Chen, D. Thota, Y. Ping, Y. Li, Theoretical and experimental insight into the effect of nitrogen doping on hydrogen evolution activity of Ni3S2 in alkaline medium. Adv. Energy Mater. 8, 1703538 (2018).

18. S. J. Deng, Y. Zhong, Y. X. Zeng, Y. D. Wang, Z. J. Yao, F. Yang, S. W. Lin, X. L. Wang, X. H. Lu, X. H. Xia, J. P. Tu, Directional construction of vertical nitrogen-doped 1T-2H MoSe2/graphene shell/core nanoflake arrays for efficient hydrogen evolution reaction. Adv. Mater. 29, 1700748 (2017).

19. M. Cabán-Acevedo, M. L. Stone, J. R. Schmidt, J. G. Thomas, Q. Ding, H.-C. Chang, M.-L. Tsai, J.-H. He, S. Jin, Efficient hydrogen evolution catalysis using ternary pyrite-type cobalt phosphosulphide. Nat. Mater. 14, 1245 (2015).

20. Y. Q. Zhang, B. Ouyang, J. Xu, S. Chen, R. S. Rawat, H. J. Fan, 3D porous hierarchical nickel-molybdenum nitrides synthesized by RF plasma as highly active and stable hydrogen-evolution-reaction electrocatalysts. Adv. Energy Mater. 6, 1600221 (2016).

21. H. J. Yan, C. G. Tian, L. Wang, A. P. Wu, M. C. Meng, L. Zhao, H. G. Fu, Phosphorus-modified tungsten nitride/reduced graphene oxide as a high-performance, non-noble-metal electrocatalyst for the hydrogen evolution reaction. Angew. Chem. Int. Ed. 54, 6325–6329 (2015).

22. B. F. Cao, G. M. Veith, J. C. Neuefeind, R. R. Adzic, P. G. Khalifah, Mixed close-packed cobalt molybdenum nitrides as non-noble metal electrocatalysts for the hydrogen evolution reaction. J. Am. Chem. Soc. 135, 19186–19192 (2013).

23. Y. P. Liu, G. T. Yu, G. D. Li, Y. H. Sun, T. Asefa, W. Chen, X. X. Zou, Coupling Mo2C with nitrogen-rich nanocarbon leads to efficient hydrogen-evolution electrocatalytic sites. Angew. Chem. Int. Ed. 127, 10902–10907 (2015).

24. Q. F. Gong, Y. Wang, Q. Hu, J. G. Zhou, R. F. Feng, P. N. Duchesne, P. Zhang, F. J. Chen, N. Han, Y. Li, C. Jin, Y. Li, S.-T. Lee, Ultrasmall and phase-pure W2C nanoparticles for efficient electrocatalytic and photoelectrochemical hydrogen evolution. Nat. Commun. 7, 13216 (2016).

25. H. B. Wu, B. Y. Xia, L. Yu, X.-Y. Yu, X. W. Lou, Porous molybdenum carbide nano-octahedrons synthesized via confined carburization in metal-organic frameworks for efficient hydrogen production. Nat. Commun. 6, 6512 (2015).

26. G. Q. Zhao, K. Rui, S. X. Dou, W. P. Sun, Heterostructures for electrochemical hydrogen evolution reaction: A Review. Adv. Funct. Mater. 28, 1803291 (2018).

27. P. Xiao, W. Chen, X. Wang, A review of phosphide-based materials for electrocatalytic hydrogen evolution. Adv. Energy Mater. 5, 1500985 (2015).

28. P. Xiao, M. A. Sk, L. Thia, X. M. Ge, R. J. Lim, J. Y. Wang, K. H. Lim, X. Wang, Molybdenum phosphide as an efficient electrocatalyst for the hydrogen evolution reaction. Energ. Environ. Sci. 7, 2624–2629 (2014).

29. S. Carenco, D. Portehault, C. Boissiere, N. Mezailles, C. Sanchez, Nanoscaled metal borides and phosphides: Recent developments and perspectives. Chem. Rev. 113, 7981–8065 (2013).

30. Y. Zhou, Y. Y. Yang, R. J. Wang, X. T. Wang, X. Y. Zhang, L. L. Qiang, W. B. Wang, Q. Wang, Z. G. Hu, Rhombic porous CoP2 nanowire arrays synthesized by alkaline etching as highly active hydrogen-evolution-reaction electrocatalysts. J. Mater. Chem. A 6, 19038–19046 (2018).

31. Y. Pan, Y. Lin, Y. J. Chen, Y. Q. Liu, C. G. Liu, Cobalt phosphide-based electrocatalysts: Synthesis and phase catalytic activity comparison for hydrogen evolution. J. Mater. Chem. A. 4, 4745–4754 (2016).

32. J. M. Wang, W. R. Yang, J. Q. Liu, CoP2 nanoparticles on reduced graphene oxide sheets as a super-efficient bifunctional electrocatalyst for full water splitting. J. Mater. Chem. A. 4, 4686–4690 (2016).

33. Z. Y. Chen, Y. Song, J. Y. Cai, X. S. Zheng, D. D. Han, Y. S. Wu, Y. P. Zang, S. W. Niu, Y. Liu, J. F. Zhu, Tailoring the d-band centers enables Co4N nanosheets to be highly active for hydrogen evolution catalysis. Angew. Chem. Int. Ed. 57, 5076–5080 (2018).

34. X. F. Xiao, C. T. He, S. L. Zhao, J. Li, W. S. Lin, Z. K. Yuan, Q. Zhang, S. Y. Wang, L. M. Dai, D. S. Yu, A general approach to cobalt-based homobimetallic phosphide ultrathin nanosheets for highly efficient oxygen evolution in alkaline media. Energ. Environ. Sci. 10, 893–899 (2017).

35. J. B. Zhou, X. J. Liu, L. Q. Zhu, J. Zhou, Y. Guan, L. Chen, S. W. Niu, J. Y. Cai, D. Sun, Y. C. Zhu, J. Du, G. M. Wang, Y. T. Qian, Deciphering the modulation essence of p bands in Co-based compounds on Li-S chemistry. Joule 2, 2681–2693 (2018).

36. Y. Tong, Y. Q. Guo, P. Z. Chen, H. F. Liu, M. X. Zhang, L. D. Zhang, W. S. Yan, W. S. Chu, C. Z. Wu, Y. Xie, Spin-state regulation of perovskite cobaltite to realize enhanced oxygen evolution activity. Chem 3, 812–821 (2017).

37. P. Z. Chen, K. Xu, Z. W. Fang, Y. Tong, J. C. Wu, X. L. Lu, X. Peng, H. Ding, C. Z. Wu, Y. Xie, Metallic Co4N porous nanowire arrays activated by surface oxidation as electrocatalysts for the oxygen evolution reaction. Angew. Chem. Int. Ed. 127, 14923–14927 (2015).

38. H. L. Fei, J. C. Dong, M. J. Arellano-Jiménez, G. L. Ye, N. D. Kim, E. L. Samuel, Z. W. Peng, Z. Zhu, F. Qin, J. M. Bao, M. J. Yacaman, P. M. Ajayan, D. L. Chen, J. M. Tour, Atomic cobalt on nitrogen-doped graphene for hydrogen generation. Nat. Commun. 6, 8668 (2015).

on August 17, 2020

http://advances.sciencemag.org/

Dow

nloaded from

Page 8: CHEMISTRY Copyright © 2020 N-induced lattice …Cai et al., ci. dv. 2020 6 : eaaw8113 3 January 2020SCIENCE ADANCES | RESEARCH ARTICLE 1 of 8 CHEMISTRY N-induced lattice contraction

Cai et al., Sci. Adv. 2020; 6 : eaaw8113 3 January 2020

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

8 of 8

39. W. Liu, E. Y. Hu, H. Jiang, Y. J. Xiang, Z. Weng, M. Li, Q. Fan, X. Q. Yu, E. I. Altman, H. L. Wang, A highly active and stable hydrogen evolution catalyst based on pyrite-structured cobalt phosphosulfide. Nat. Commun. 7, 10771 (2016).

40. A. Kahyarian, B. Brown, S. Nesic, Mechanism of the hydrogen evolution reaction in mildly acidic environments on gold. J. Electrochem. Soc. 164, H365–H374 (2017).

41. S. Fletcher, Tafel slopes from first principles. J. Solid State Electrochem. 13, 537–549 (2009). 42. Y. S. Wu, X. J. Liu, D. D. Han, X. Y. Song, L. Shi, Y. Song, S. W. Niu, Y. F. Xie, J. Y. Cai, S. Y. Wu,

J. Kang, J. B. Zhou, Z. Y. Chen, X. S. Zheng, X. H. Xiao, G. M. Wang, Electron density modulation of NiCo2S4 nanowires by nitrogen incorporation for highly efficient hydrogen evolution catalysis. Nat. Commun. 9, 1425 (2018).

43. J. R. Kitchin, J. K. Nørskov, M. A. Barteau, J. Chen, Role of strain and ligand effects in the modification of the electronic and chemical properties of bimetallic surfaces. Phys. Rev. Lett. 93, 156801 (2004).

44. B. Hammer, J. K. Nørskov, Theoretical surface science and catalysis—Calculations and concepts. Adv. Catal. 45, 71–129 (2000).

45. Y. H. Liang, Q. Liu, A. M. Asiri, X. P. Sun, Y. L. Luo, Self-supported FeP nanorod arrays: A cost-effective 3D hydrogen evolution cathode with high catalytic activity. ACS Catal. 4, 4065–4069 (2014).

46. J. M. Wang, X. Ma, F. L. Qu, A. M. Asiri, X. P. Sun, Fe-doped Ni2P nanosheet array for high-efficiency electrochemical water oxidation. Inorg. Chem. 56, 1041–1044 (2017).

47. J. Xie, L. Gao, H. Jiang, X. Zhang, F. Lei, P. Hao, B. Tang, Y. Xie, Platinum nanocrystals decorated on defect-rich MoS2 nanosheets for pH-universal hydrogen evolution reaction. Cryst. Growth Desi. 19, 60–65 (2019).

48. R. Zhang, X. Wang, S. Yu, T. Wen, X. Zhu, F. Yang, X. Sun, X. Wang, W. Hu, Ternary NiCo2Px nanowires as pH-universal electrocatalysts for highly efficient hydrogen evolution reaction. Adv. Mater. 29, 1605502 (2017).

49. S. J. Clark, M. D. Segall, C. J. Pickard, P. J. Hasnip, M. I. Probert, K. Refson, M. C. Payne, First principles methods using CASTEP. Z. Krist-Cryst. Mater. 220, 567–570 (2005).

50. J. K. Nørskov, T. Bligaard, A. Logadottir, J. Kitchin, J. G. Chen, S. Pandelov, U. Stimming, Trends in the exchange current for hydrogen evolution. J. Electrochem. Soc. 152, J23–J26 (2005).

Acknowledgments: We acknowledge the Beamline BL14W1 at Shanghai Synchrotron Radiation Facility for XAS characterizations. The numerical calculations in this paper were performed on the supercomputing system in the Supercomputing Center of University of Science and Technology of China. Funding: This work was supported by the National Key Research and Development Program of China (2017YFA0206703), the Natural Science Fund of China (nos. 21771169, 11722543, and 11505187), the Fundamental Research Funds for the Central Universities (WK2060190074, WK2060190081, and WK2310000066), and the Recruitment Program of Global Expert. Author contributions: G.W., Y.Q., X.L., and Y.Lin designed and supervised the project. J.C. and Y.S. conducted the project. X.Z., Y.Z., and Y.W. provided the XAFS and XPS measurements and analysis. S.N., Y.X., and Y.Liu helped in the material synthesis. G.W., X.L., J.C., and Y.S. wrote and revised the manuscript. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials. Additional data related to this paper may be requested from the authors.

Submitted 27 January 2019Accepted 7 November 2019Published 3 January 202010.1126/sciadv.aaw8113

Citation: J. Cai, Y. Song, Y. Zang, S. Niu, Y. Wu, Y. Xie, X. Zheng, Y. Liu, Y. Lin, X. Liu, G. Wang, Y. Qian, N-induced lattice contraction generally boosts the hydrogen evolution catalysis of P-rich metal phosphides. Sci. Adv. 6, eaaw8113 (2020).

on August 17, 2020

http://advances.sciencemag.org/

Dow

nloaded from

Page 9: CHEMISTRY Copyright © 2020 N-induced lattice …Cai et al., ci. dv. 2020 6 : eaaw8113 3 January 2020SCIENCE ADANCES | RESEARCH ARTICLE 1 of 8 CHEMISTRY N-induced lattice contraction

metal phosphidesN-induced lattice contraction generally boosts the hydrogen evolution catalysis of P-rich

Gongming Wang and Yitai QianJinyan Cai, Yao Song, Yipeng Zang, Shuwen Niu, Yishang Wu, Yufang Xie, Xusheng Zheng, Yun Liu, Yue Lin, Xiaojing Liu,

DOI: 10.1126/sciadv.aaw8113 (1), eaaw8113.6Sci Adv 

ARTICLE TOOLS http://advances.sciencemag.org/content/6/1/eaaw8113

MATERIALSSUPPLEMENTARY http://advances.sciencemag.org/content/suppl/2019/12/20/6.1.eaaw8113.DC1

REFERENCES

http://advances.sciencemag.org/content/6/1/eaaw8113#BIBLThis article cites 50 articles, 2 of which you can access for free

PERMISSIONS http://www.sciencemag.org/help/reprints-and-permissions

Terms of ServiceUse of this article is subject to the

is a registered trademark of AAAS.Science AdvancesYork Avenue NW, Washington, DC 20005. The title (ISSN 2375-2548) is published by the American Association for the Advancement of Science, 1200 NewScience Advances

License 4.0 (CC BY-NC).Science. No claim to original U.S. Government Works. Distributed under a Creative Commons Attribution NonCommercial Copyright © 2020 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of

on August 17, 2020

http://advances.sciencemag.org/

Dow

nloaded from