Electronic Structure Investigations. A novel model with Ru-Ni3N/NiO three-phase heterogeneous interface was firstly constructed together with the Ru-Ni3N, Ru-NiO, and Ni3N/NiO models (Supplementary Fig. 1). The bonding and anti-bonding orbitals distributions near the Fermi levels (EF) have shown different behaviors for the heterostructure systems (Fig. 1a-c). Within Ru-Ni3N, it is noted that the anti-bonding orbitals are dominated by the anchoring Ru NPs while the bonding orbitals mainly locate in Ni sites with limited contributions from Ru NPs. For Ru-NiO, we notice the structural instability has evidently increased due to the
evident distortion in NiO. Although the bonding orbitals become more electron-rich in Ru NPs, the lower lattice stability may hinder the electrocatalysis performances. In comparison, the Ru-Ni3N/NiO heterostructure displays a highly electron-rich feature on the surface and Ru NPs to guarantee efficient electron transfer with a stable structure. The strong orbital coupling between bonding and anti-bonding orbitals results in superior electroactivity for both oxidation and reduction reactions. The detailed electronic structures are demonstrated by the partial density of states (PDOS) for these three heterostructures Ru-Ni3N, Ru-NiO, and Ru-Ni3N/NiO systems (Fig. 1d-f). Notably, all Ni-3d orbitals show evident a sharp peak near the EF, which play as the dominant sites for proton binding during the HER. Meanwhile, Ru-4d orbitals all exhibit a broad peak with eg-t2g splitting, which plays as the main active site for water-dissociation under the alkaline environment. Notably, N-2p and O-2p orbitals are both located at the deeper position, which act as the electron reservoir during electrocatalysis. For Ru-Ni3N/NiO, it is noted that orbital coupling between O and N further strengthens the site-to-site electron transfer within the heterostructure. The site-dependent PDOSs further reveal the electronic modulation induced by the heterostructure (Fig. 1g). For Ru NPs, we notice that from the interface with Ni3N/NiO to the NP surface, the eg-t2g splitting is strongly alleviated from 4.70 eV to 1.24 eV, indicating much alleviated barriers for electron transfer, which is able to accelerate the water dissociation process. Meanwhile, the Ni-3d orbitals display different behaviors in Ni3N and NiO components (Fig. 1h). Within the Ni3N, the formation of the interface with NiO has led to the upshifting of the Ni-3d orbitals with improved electroactivity. The introduction of Ru NPs on the surface further upshifts the Ni-3d orbitals to guarantee efficient electron transfer. In comparison, the Ni-3d orbitals in NiO experience a volcano trend (Fig. 1i). From the bulk to the surface, Ni-3d orbitals exhibit a gradual upshifting trend. For the Ni sites at the interface with Ni3N and Ru NPs, their 3d orbitals show slight downshifting. Such a volcano trend indicates that the formation of heterostructures enables subtle modulations of the electroactivity. The d-band center of Ru in Ru-Ni3N/NiO balances the electroactivity of water dissociation and binding of OH* (Fig. 1j). In contrast, the d-band center of Ni sites shows a gradual downshifting trend from Ru-Ni3N to Ru-Ni3N/NiO, which optimizes the overbinding effect of protons. Thus, it can be inferred that Ru-Ni3N/NiO may have very excellent electrocatalytic EWS activity.
Synthesis and morphological characterizations. To further validate experimentally the superior activity of Ru NPs on Ni3N/NiO substrates with heterogeneous structures, we designed and prepared the obtained Ru- Ni3N/NiO catalysts. The synthetic route of Ru-Ni3N/NiO electrocatalysts was exhibited in Fig. 2a (see experimental details in the Methods). First, the ultrathin Ni(OH)2 nanosheet arrays were in-situ grown on the
3D porous Ni foam (NF) through simple acid corrosion engineering. After annealing in NH3, the Ni(OH)2 nanosheets were transformed into Ni3N/NiO heterostructured nanosheets. Subsequently, Ru nanoparticles were formed onto the heterostructured Ni3N/NiO nanosheets by a spontaneous redox reaction (2Ru3+ + 3Ni → 2Ru + 3Ni2+) in RuCl3 solution at room temperature12, 32.
The scanning electron microscopy (SEM) characterization indicates the successful formation of the uniform Ni(OH)2 ultrathin nanosheet arrays on the NF surface with a height of about 1 ~ 2 µm (Fig. 2b and Supplementary Fig. 2) and a thickness of ~ 1.5 nm (Fig. 2c). After low temperature annealing in NH3 atmosphere, the obtained Ni3N/NiO material maintains the nanosheet arrays morphology yet with a rougher and defect-rich surface (Supplementary Figs. 3 and 4). As can be observed in Supplementary Fig. 5, the successful synthesis of Ni3N/NiO heterostructured nanosheets was confirmed by the high-resolution transmission electron microscopy (HRTEM) characterization. And monodisperse Ru nanoparticles can be found on the surface of the ultrathin Ni3N/NiO nanosheet (Fig. 2d, e and Supplementary Fig. 6). Meanwhile, the morphology of the Ru-Ni3N/NiO composite is strictly affected by the immersion time, where too long or too short impregnation time is not conducive to the formation of Ru-Ni3N/NiO (Supplementary Fig. 7). The HRTEM image of Ru-Ni3N/NiO (Fig. 2f and Supplementary Fig. 8) shows that a three-phase heterogeneous interface (dotted line) formed in the resulting Ru-Ni3N/NiO catalyst and the lattice fringes are mainly ascribed to the (101) plane of Ru, the (002) plane of Ni3N, and the (111), (200) planes of NiO. The energy-dispersive spectra (EDS) analysis results further indicate that the nanoparticles loaded on the Ni3N/NiO nanosheets are Ru (Fig. 2g). The high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) and the corresponding EDS elemental mapping images demonstrate the successful introducing of Ni, O, N and Ru elements in Ru-Ni3N/NiO (Fig. 2h).
Structure characterization. The composition and surface valence of the obtained catalysts were investigated by X-ray diffraction (XRD) and X-ray photoelectron spectroscopy (XPS). The XRD peaks of Ru-Ni3N/NiO in Fig. 3a are corresponded well with Ru (JCPDS 06-0663), Ni3N (JCPDS 10–0280), and NiO (JCPDS 44-1159), confirming the formation of these three phases in Ru-Ni3N/NiO. The XPS survey spectra verified the elemental composition and content of Ru-Ni3N/NiO and Ni3N/NiO (Fig. 3b and Supplementary Table 1). As
indicated in the high-resolution Ru 3p spectrum of Ru-Ni3N/NiO (Fig. 3c), two peaks with binding energies (BEs) of 463.38 eV and 484.38 eV are corresponded to Ru 3p3/2 and Ru 3p1/2 of metallic Ru, respectively, demonstrating the existence of metallic Ru (0)33, 34. As shown in Fig. 3d, the high-resolution Ni 2p spectra of Ru-Ni3N/NiO and Ni3N/NiO display the typical peaks of Ni 2p3/2 (856.08 eV) and Ni 2p1/2 (873.48 eV)12. In comparison of Ni3N/NiO, Ni peaks of Ru-Ni3N/NiO shifted to a higher energy level by 0.2 eV, demonstrating the strong electronic MSI between metallic Ru and Ni3N/NiO carrier. And this MSI result can be further verified by the positive shifts of 0.2 eV in high-resolution N 1s spectra of Ni3N/NiO and Ru-Ni3N/NiO (Fig. 3e). The high-resolution O 1s spectrum of Ni3N/NiO was assigned to three peaks, which are metal-oxygen bonds (ONi−O), low coordination oxygen ions (OOL), and water adsorbed on the material surface (OH−OH)35, 36. Among them, the OOL peak centered at 531.5 eV can be attributed to the presence of defective sites. And the OOL peak ratio would greatly increase to 90.5% in the Ru-Ni3N/NiO sample, indicating the formation of much more defect sites in Ru-Ni3N/NiO(Fig. 3f)37.
Electrochemical HER. The HER properties of different samples were investigated in N2-saturated 1.0 M KOH solution, and electrochemical tests were performed in a three-electrode system. The iR-compensated linear scanning voltammogram (LSV) curves for HER over Ru-Ni3N/NiO, Ru-Ni3N, Ru-NiO, Ni3N/NiO, commercial Pt/C and NF can be observed in Fig. 4a. Among them, the target Ru-Ni3N/NiO catalyst exhibits the best HER performance with overpotentials of 15 mV, 46 mV, and 63 mV at current densities of 20 mA cm− 2, 50 mA cm− 2 and 100 mA cm− 2, respectively. And compared to most of the reported HER catalysts, Ru-Ni3N/NiO exhibit excellent activity at low potential (Supplementary Table 2). More importantly, Ru-Ni3N/NiO also shows surprising performance at industrial current densities, with overpotentials of only 121 mV and 190 mV at current densities of 500 mA cm− 2 and 1000 mA cm− 2, demonstrating its potential for industrial applications. Furthermore, the overpotential vs. Tafel slope (Fig. 4b and Supplementary Fig. 10) reveal Ru-Ni3N/NiO performs Pt-like HER kinetics with a Tafel slope of 45.1 mV dec− 1, which is much better than other contrast samples. (Ru-Ni3N (47.8 mV dec− 1), Ru-NiO (59.9 mV dec− 1), Ni3N/NiO (98.8 mV dec− 1), commercial Pt/C (51.85 mV dec− 1) and NF (118 mV dec− 1) (Fig. 4b and Supplementary Table 3).
Double-layer capacitance (Cdl) was usually applied to evaluate the electrochemical specific area (ECSA) of different modified electrodes (Supplementary Fig. 11). The Ru-Ni3N/NiO electrode owns largest Cdl of 311.7 mF cm− 2 among all contrast samples (Supplementary Fig. 12), indicating more electrochemical active
sites can be exposed in Ru-Ni3N/NiO. Meanwhile, the lowest charge transfer resistance (Rct) of Ru-Ni3N/NiO was gained by the electrochemical impedance spectroscopy (EIS) tests, revealing a desirable electron transport ability in Ru-Ni3N/NiO (Supplementary Fig. 13 and Supplementary Table 4). Furthermore, extremely small contact angle of Ru-Ni3N/NiO (~ 0°) can greatly promote rapid penetration and mass transfer of the electrolyte in the Ru-Ni3N/NiO electrode (Supplementary Fig. 14).
Subsequently, the electrochemical long-time stability of Ru-Ni3N/NiO was investigated by chronopotentiometry (CP) as well as accelerated degradation test (ADT). And the polarization curve of Ru-Ni3N/NiO hardly shifts after 15,000 CV cycles (Fig. 4c). As for the CP test, Ru-Ni3N/NiO can operate better stability than the Pt/C catalyst at different current density (100 mA cm− 2 and 500 mA cm− 2) (Supplementary Figs. 15–17). After the durability test, the morphology and valence state of Ru-Ni3N/NiO with no change could be observed by SEM and XPS (Supplementary Figs. 18–19), further demonstrating its superior robustness. These results indicated that large ECSA and rapid charge/mass transfer capability of Ru-Ni3N/NiO synergistically prompt it enhanced HER catalytic activity and stability.
Electrochemical OER. Meanwhile, the Ru-Ni3N/NiO catalyst exhibits better OER performance with a smaller over potential at 100 mA cm− 2 current density (\(\eta\)100 = 261 mV), than Ru-Ni3N (\(\eta\)100 = 321 mV), Ru-NiO (\(\eta\)100 = 337 mV), Ni3N/NiO (\(\eta\)100 = 379 mV), RuO2 (\(\eta\)100 = 404 mV), and NF (\(\eta\)100 = 451 mV) samples (Fig. 4d). And a high current density up to 1000 mA cm− 2 can be achieved at an overpotential of 385 mV for Ru-Ni3N/NiO due to the unique structural advantage and the enhanced MSI. More significantly, Ru-Ni3N/NiO has the lowest tafel slope compared to all the comparison samples synthesized (Fig. 4e and Supplementary Fig. 10, Supplementary Table 3), which can be comparable to the reported state-of-art OER catalysts (Supplementary Table 5). The Ru-Ni3N/NiO catalyst also shows excellent catalytic stability for OER with no significant change after 15000 CV cycles and CP stability test (> 100 h) (Fig. 4f and Supplementary Figs. 20–21). After the OER long-term durability test, the original nanosheet array structure was maintained, and slight of the Ru nanoparticles were oxidized to Ru4+, thus promoting the OER activity (Supplementary Figs. 22–23).
Reaction Trends by DFT Calculations. DFT calculations were used to further reveal the intrinsic relationship between the modulated electronic structure and the superior performances of Ru-Ni3N/NiO under alkaline conditions. First, the reaction energy of HER has been compared (Fig. 4g). Notably, the initial adsorption of water and protons are both energetically favored for all three heterostructure systems. The dissociation of water is the rate-determining step (RDS) of the HER, where the energy barrier is 0.83, 0.73, and 0.60 eV for Ru-Ni3N, Ru-NiO, and Ru-Ni3N/NiO, respectively. Meanwhile, the reaction energy change of Ru-Ni3N and Ru-NiO are similar, which are smaller than that of the Ru-Ni3N/NiO, supporting the superior performances of Ru-Ni3N/NiO for the HER. Then, the reaction energy of OER is also compared, where the RDS is the conversion from O* to OOH* with the largest energy barrier (Fig. 4h). Ru-Ni3N/NiO displays the smallest barrier of 1.43 eV for the RDS, leading to the highest OER performance. With the applied equilibrium potential (U = 1.23 V), the overpotential has been estimated to be 0.20 V for Ru-Ni3N/NiO, which is much alleviated than that of Ru-Ni3N, Ru-NiO, supporting the improved OER performances (Fig. 4i). The introduction of Ru nanoparticles on Ni3N/NiO heterojunction carriers and the formation of a unique heterostructure resulted in superior electrocatalysis as the electronic structures were optimized and the energy barrier of RDS for HER and OER has been significantly alleviated.
Performance in water electrolyzers device. Encouraged by the excellent catalytic activity of Ru-Ni3N/NiO for both HER and OER, a two-electrode system was assembled using the Ru-Ni3N/NiO as both cathode and anode (Ru-Ni3N/NiO(+,-)) for overall water splitting. As shown in Fig. 5a, only a small cell voltage of 1.74 V is needed to achieve an industrial current density (1000 mA cm− 2) in this Ru-Ni3N/NiO(+,-)-based overall water splitting device. Meanwhile, the Ru-Ni3N/NiO(+,-) electrolyzer powered by a 1.5 V battery can work well in 1.0 M KOH with significant bubble release (Fig. 5b). Moreover, the Ru-Ni3N/NiO(+,-) electrolyzer showed surprising durability, maintaining a cell voltage of 1.74 V for over 1000 h at industrial current density of 500 mA cm− 2 with almost no significant degradation, further confirming the excellent stability of the catalyst (Fig. 5c).
To explore their practical application potential, the Ru-Ni3N/NiO catalyst were tested in the alkaline seawater (0.5 M KOH + seawater), which display comparable HER and OER activity compare to the results in 1.0 M KOH solution (Supplementary Figs. 24–25). More importantly, Ru-Ni3N/NiO(+,-) electrolyzer in alkaline SW and alkaline urea (1.0 M KOH + 0.33 M urea) could also achieve a good electrochemical performance (with a potential of only 1.75 V and 1.53 V at 500 mA cm− 2 ) and long-time durability (@ 500 mA cm− 2, 50 h), respectively (Supplementary Figs. 26–27). All these results prove that the obtained Ru-Ni3N/NiO catalyst is one of the best reported bifunctional catalysts for the EWS process (Fig. 5d and Supplementary Table 6), and in particular it exhibits excellent performance at a high current densiy (500 mA cm− 2) (Supplementary Table 7).