With conventional metal oxide or sulfide photocatalysts18,22, the surface active site tends to bond weakly with adsorbed N2 through hybridization between the 2p and 3d orbital (top panel of Fig. 1a), which leads to a series of reactive intermediates23,24. However, compared with the highly stable N ≡ N triplet bonds, the relatively weak M-N bonds (where M represents the metal site) needs much more energy to satisfy NH3 production. One can imagine that if N2 could be simultaneously affected by metal sites and passivated H*, this would give rise to a more stable configuration of M-HN-NH-M intermediates and high NRR efficiency (middle panel of Fig. 1a), owing to integration reaction between surface hydrogenation in HER and N2 activation in NRR. Interestingly, it becomes difficult once again to break N ≡ N triplet bond to form NH3, supposing that passivated H* from HER is released as H2 (bottom panel of Fig. 1a). Therefore, the rate-limiting activation energy barrier in NRR may be significantly decreased through a conversion in reaction pathway, after considering dual photocatalytic integration.
To better understand this fundamental mechanism, a crystal-field theory to describe the molecular orbital energy diagram is proposed in Fig. 1b. For transition metal catalysts, their reaction activity strongly depends on the combination of empty and occupied 3d orbitals. To realize the molecular chemisorption, the metal sites need to empty 3d orbitals to accept the lone-pair electrons of N2. On the contrary, to enhance the activity of N-M bonds, the metal atoms should have separate 3d electrons that can donate into the antibonding orbital to weak the N ≡ N triplet bond. Generally speaking, “acceptance-donation” of electrons is the essential interaction between the metal site and N2 (left panel of Fig. 1b)25, in which regulation of empty and occupied 3d orbital plays a critical role in enhancing catalytic activity. Take Fe site for example, the filling of this eg-like and t2g-like states depends on the number of 3d electrons and the spin configuration. The corresponding t2g occupancies before and after hydrogenation for Fe3+ is 3 and 4, respectively (middle panel of Fig. 1b). When N2 adsorbs onto the hydrogenated Fe sites, the relevant adsorbate molecules interact with vertically oriented eg-like orbitals26, which allows for stronger overlap with t2g-like states (right panel of Fig. 1b). Therefore, the hydrogenated Fe atom can serve as an active center for driving the “acceptance-donation” process, resulting in an efficient reduction of N2.
Co0.5Fe0.5In2S4 is an earth-abundant and non-toxic quaternary sulfide27 that has desirable electron-hole separation and controllable catalytic activity in HER and NRR, which suggests it should be capable of integrating reaction pathway. After the analysis about electronic features and its chemical structure, we reasoned that it should be a promising candidate to integrate HER and NRR by manipulating catalytic activity of reactive site and life time of photo-generated carrier simultaneously. Therefore, we initially fabricated Co0.5Fe0.5In2S4 ultrathin nanoflowers by a citric-assisted solvothermal procedure with cobaltous nitrate, iron 2,4-pentanedionate, indium chloride and thioacetamide as the Co, Fe, In and S sources, respectively. The scanning electron microscopy (SEM) and transmission electron microscopy (TEM) images in Fig. 2a display that the Co0.5Fe0.5In2S4 nanoflowers are vertically aligned with three-dimensional porous textures, indicating distinguished capture of sunlight. The atomic force microscopy (AFM) in Fig. 2b discloses that the exfoliated self-assembly nanosheet is of 8 nm in thickness and 0.5 µm in width. The high-resolution TEM (HR-TEM) images in Fig. 2c and selected diffraction pattern in the inset confirm that the Co0.5Fe0.5In2S4 nanoflowers have a typical polycrystalline structure, with a resolved 0.32 nm lattice fringe corresponding to (311) planes. The corresponding element mapping in bottom panel confirms the uniform distribution of Co, Fe, In and S elements, and their atomic ratio was found to 0.5:0.5:2.1:3.9 associated with energy-dispersive spectroscopy (Supplementary Fig. 1). As a comparison, the X-ray diffraction (XRD) patterns of ternary FeIn2S4 and CoIn2S4 sample are also provided, which matches well with JCPDS card No. 80–0609 and No.65-7271. Remarkably, the XRD pattern of Co0.5Fe0.5In2S4 sample has similar fingerprint to cubic CoIn2S4 phase, showing that expected Fe implantation cannot change the original crystal structure. The slight difference in XRD fingerprint can be attributed to the alloy-induced changes in geometrical structure factor.
X-ray absorption fine structure (XAFS) was conducted to investigate the atomic-hybridized influence towards the electronic structure. The energies of X-ray absorption near edge structure (XANES) for Fe K-edges and Co K-edges in CoIn2S4, FeIn2S4, Co0.5Fe0.5In2S4 are positively shifted compared with those of Fe/Co foils (Fig. 3a and b), showing the oxidized states of Fe/Co. The XANES spectra of Co and Fe in Co0.5Fe0.5In2S4 display similar patterns to CoIn2S4 and FeIn2S4, but the peak intensities and positions are slightly changed due to the difference in radial charge distribution for Co-3d and Fe-3d atomic orbitals (Supplementary Fig. 2). Fourier transform analysis of the corresponding K-edge extended EXANES was then used to verify their distinct local structures. As displayed in Fig. 3c, the Fourier transform spectrum of FeIn2S4 centred at 1.79 Å corresponds to the nearest Fe-S coordination, which also occurs at Co0.5Fe0.5In2S4 sample. However, the Fe-S peak intensity obtained from Co0.5Fe0.5In2S4 is weak and this peak position shifts to a large bond length (~ 1.81 Å). A similar change associated with the Co-S coordination is also observed in Fig. 3d. The aforementioned difference suggests that the Co-S and Fe-S shell in Co0.5Fe0.5In2S4 is distorted comparing to pristine CoIn2S4 and FeIn2S4 and its structural homogeneity is changed by hybridization of atomic orbitals. Moreover, the curve fitting of coordination environment was conducted using IFEFFIT software package to acquire the quantitative structural parameters28, and the corresponding results (Supplementary Fig. 3) are listed in Supplementary Table 1. It was found that the coordination numbers of CoIn2S4 (Co: 4.6 ± 0.7), FeIn2S4 (Fe: 3.7 ± 0.5) and Co0.5Fe0.5In2S4 (Co: 4.1 ± 0.5; Fe: 3.6 ± 0.7) are very similar, indicating that they have consistent space symmetry. Compared to Co0.5Fe0.5In2S4, the Fe-S bond length of FeIn2S4 is reduced by 0.02 Å, while the Co-S bond length of CoIn2S4 was increased by 0.02 Å. This indicates that the Fe and Co atoms are equivalently distributed in Co0.5Fe0.5In2S4 structure. However, the electronegativity difference between Fe (1.83) and Co (1.88) make their chemical state display a discrepancy, due to different valence electron ability in transferring to surrounding sulfur site, which can be reflected by XPS analysis in Supplementary Fig. 4. As a result, the valence state of Fe and Co ion lineally depend on alloy compositions (Supplementary Fig. 2), implying that nonequilibrium charge site is fabricated to regulate photo-excited carrier life time. This is because that the carrier transfer on the catalyst surface strongly depends on atomic-hybridized-induced electrostatic potential difference29.
As Co0.5Fe0.5In2S4 surface shown in Fig. 3e, the first layer has a lower electrostatic potential than the second one due to the asymmetric charge density for Fe and Co site (as shown in the right panel), leading to an instinct dipole moment for generating a built-in electric field between two adjacent atomic layers (marked by Δδ). Interestingly, this electrostatic potential difference cannot occur at pristine CoIn2S4 and FeIn2S4 surface, showing that the electronic state can be regulated by Co-Fe alloy strategy. The electrostatic potential difference as a function of alloy composition in bottom of Fig. 3f disclose that alloys with 0.5:0.5 Fe/Co ratio display the optimal performance, which can be intuitively reflected by Bader charge analysis shown in the inset.
Subsequently, we experimentally confirm the atomic-hybridized contribution to extend carrier life time using a transient absorption technology. The orbital-resolved band of Co0.5Fe0.5In2S4 in Fig. 4a displays an indirect semiconductor characteristic, similar with that of CoIn2S4 and FeIn2S4 in Supplementary Fig. 5. Interestingly, we find that the valence bands at G point are mainly contributed by Fe-d orbitals, meanwhile the bands of Co-d orbitals appear at deep level energy. This multi-level system provides a feasibility in regulating photo-excited carrier dynamics, as the discussion about electrostatic potential difference in Fig. 3e and f. In order to confirm this assertion, for a probe photon from 1.20 to 1.65 eV and a time delay of 0.00 to 2.00 ps, a three-dimensional picture of reflective transient absorption (TA) spectroscopic measurement with a pump fluence of 400 µJ/cm2 at 1.40 eV is conducted and shown in Fig. 4b. The most typical fingerprint in Co0.5Fe0.5In2S4 is the strong negative TA within the first 500 fs, which is located around 1.42 eV. To better understand this physical origin, the TA dynamics for a time delay as long as 8.00 ps at 1.34, 1.36, 1.44 and 1.66 eV are acquired as a comparison in Fig. 4c. Noticeably, an oscillation appear at 1.44 eV, and the signal reverts to positive at 0.69 ps and then reach a positive maximum at 0.85 ps. It is interesting to note that the TA signals remain negative at 1.66 eV, meanwhile it keeps positive at 1.34 and 1.36 eV over the whole dynamics.
For better understanding, we aligned the total density of state (TDOS) of Co0.5Fe0.5In2S4 with the electronic transitions disclosed by the TA spectra, as displayed in the left panel of Fig. 4d. The black open circle curve is a fit using multiple Gaussian functions, which are designated by the different shaded area and named as bands D0, D1, E0, E1, respectively. Associated band analysis in Fig. 4a, the ground-state bands only can be attributed to the band D1 (Fe-3d at G point) instead of band D0 that originates from Fe-3d orbital at L point. The lowest excited state is located at the lower edge of band E0. This assignment can be confirmed by the optical transition moment ρ2, in which the transition probability at G point from band D1 to E0 becomes dominant, as indicated by the bottom panel of Fig. 4a. Associated with these energy levels, we conclude the possible electronic transitions involved in the pump-probe process in the right panel of Fig. 4d. Therefore, the TA signal includes bleaching of the ground-state absorption at D1 (❸), absorption by exactions at E0 (❹), two-photon absorption (2PA) through absorbing one pump and one probe photon from the ground state D1 (❺) and absorption by excitation at higher-lying excited state E1 (❻). Then the oscillatory feature in the TA dynamics at 1.42 eV (Fig. 4e) can be explained due to existence of higher-lying E1 state. This is because that bleaching of ground state absorption (negative signal) plays a decisive role in the TA dynamics at the arrival of the pump pulses, in which single-photon excitation (1PE) (❶) and two-photon excitation (2PE) pump (❷) the D1 electron to populate E1 and E0, respectively. Subsequently, after considering the probe contribution, the exciton absorption at E1 and E0 state (allowed electronic transition in ① and ② process) plays a dominant role, which leads to a positive TA signal. When this exciton absorption survives other competitive bleaching dynamics, the TA signal exhibit a turn from falling to rising dynamics at 0.44 ps and finally becomes positive signal at 0.69 ps. Because of the shorter life time of higher-lying metastable E1 state, the long-lived bleaching dynamics dominate the electronic transition process again (③), converging towards a negative TA signal after 1.87 ps.
The exciton behavior of E1 also can be confirmed by the dynamics of pump fluence dependence, as shown in Fig. 4 f. For a delay of 1.10 ps, owing to the existence of metastable state E1, the relative TA intensity is increased with third-order dependence on the pump fluence. This process includes exciton absorption at E0 and E1state, which can be defined as here Ip, Id denote the intensities of pump and probe pulses. When higher-lying metastable state E1 disappears at 2.80 ps, the relation between pump fluence and TA signal is changed as As a critical criterion, compared to CoIn2S4 and FeIn2S4, the Co0.5Fe0.5In2S4 sample in Supplementary Fig. 6 displays a typical three-stage evolution dynamic behavior and the photo-excited carrier life time is also obviously extended by atomic-hybridized-generated higher-lying metastable state.
The photocatalytic NRR performances of different samples were conducted to test our proposed dual-photocatalytic integration, and detailed descriptions about experimental measurement and conditions are provided in the Methods. As shown in Fig. 5a, the pristine CoIn2S4 and FeIn2S4 display NH3 evolution rates of 21.8 µmol g− 1 h− 1 and 38.3 µmol g− 1 h− 1, which is obviously higher than that of commercial TiO2. Their photolysis yields are monotonically increased with photolysis time. Remarkably, average NH3 evolution rate of Co0.5Fe0.5In2S4 is enhanced to 85.8 µmol g− 1 h− 1, 3.9 (2.2) times higher than that of pristine CoIn2S4 and FeIn2S4 (Supplementary Figs. 7–10). Meanwhile, the apparent quantum efficiency (AQE) of Co0.5Fe0.5In2S4 was estimated to be 1.65% at 350 nm, 1.09% at 400 nm and 0.12% at 450 nm (Supplementary Fig. 11). After four successive photocatalytic cycles over 12 hours in Supplementary Fig. 12, the Co0.5Fe0.5In2S4 also keeps excellent structural stability and photocatalytic performance. Compared to other photocatalysts under similar conditions as a potential application, the Co0.5Fe0.5In2S4 catalyst exhibits a state-of-the-art N2 reduction performance, as shown in Supplementary Table 2. In addition, a series of isotopic labelling and verified experiments are also conducted in Supplementary Figs. 13–14, which indicates that the evolved NH3 indeed originates from the photoreduction of N2. Notably, the high selectivity of Co0.5Fe0.5In2S4 for N2 photoreduction into ammonia also can be verified, because no apparent hydrazine was detected in the product solutions (Supplementary Fig. 15)
To confirm the photocatalytic integration between HER and NRR, we conducted in situ Raman spectroscopy to detect the reaction intermediates in the N2 reduction process. After light irradiation, a new Raman peak at 1028.53 cm− 1 for Co0.5 Fe0.5In2S4 in water solution and Ar atmosphere is detected, and this peak subsequently disappear after removing light source. Naturally, this peak only can be attributed to passivated H* group on metal site due to H2O splitting, which also can be confirmed by the pH-dependent Raman behavior, as shown in Supplementary Fig. 16. When N2 is introduced into this reaction process, the Raman intensity at 1028.53 cm− 1 is depressed accordingly, while a stronger Raman peak is detected at 960.39 cm− 1. This Raman behavior can be assigned to symmetric stretching vibration of *N2H2 group on Co0.5Fe0.5In2S4 surface30–33, which well coincides with the calculated results of 975.73 cm− 1. This interdependent spectral behavior indicates passivated H* group is a crucial intermediate during N2 reduction to NH3. In other words, dual photocatalytic integration from HER to NRR is realized and this strategy leads to an excellent NH3 evolution efficiency.
To detect the underlying physical mechanism, Fig. 5c shows the H* free energy (ΔGH) (left axis) and N2H2 dissociation energy (right axis) as a function of charge-transfer energy (energy difference between occupied metal 3d and unoccupied N 2p-band center or H 1 s-band center, top panel in Fig. 5c) for different metal site of (001) surface in different alloy compositions. The charge-transfer energy is strongly dependent of type of metal site and its coordination environment, displaying a linear dependence relation with catalytic activity. The Co, Fe, Co(1), Co(2), Fe(1), Fe(2), In-Co, In-Fe, In-CoFe represents different reactive sites for HER and CoFe, Fe(1&2), Co(1&2) represents different reactive sites for NRR, details shown in Supplementary Fig. 17. The calculation discloses that Co and Fe site (marked by grey Co and Fe) at Co0.5Fe0.5In2S4 surface have smaller ΔGH than corresponding In site (marked by grey In-CoFe), indicating the HER prefers to occur at Co and Fe site. Additionally, the HER activity of Co0.5Fe0.5In2S4 is obvious excellent than that of CoIn2S4 (Co(1), Co(2) and In-Co) and FeIn2S4 (Fe(1), Fe(2) and In-Fe). It is generally accepted that the ΔGH can be used to evaluate the HER activity. The optimal value of ΔGH is zero34, where hydrogen is bound neither too strongly nor too weakly. The negative ΔGH (-0.10 eV) in Co0.5Fe0.5In2S4 implies a not outstanding HER performance due to an existence of rate-limiting potential. From another perspective, this not excellent HER performance provides a feasibility for integrating following NRR, which can be confirmed by changes in hydrogen production in N2 and Ar atmosphere (Supplementary Fig. 18). The detailed reaction pathway for NRR is shown in Supplementary Fig. 19 and the N2H2 dissociation energy (rate-limiting step) is extracted and described as a function of charge-transfer energy in Fig. 5c. The formation of metal-nitrogen bond has a mixed ionic-covalent character because of the energetic similarity (covalency) and spatial overlap (hybridization) between metal 3d orbitals and N 2p states, which plays a critical role in NRR. In alloying design, replacing the Co site with a more electronegative Fe atom to form Co0.5Fe0.5In2S4, the enhancing valence state of Fe site (Supplementary Fig. 2) increases the Fe-N covalency and hybridization, resulting in a lower N2H2 dissociation energy (marked by red CoFe). From the understanding of reaction pathway in Supplementary Fig. 19, the integration of HER and NRR is a good choice to enhance the photocatalytic performance.
In order to further confirm this dual-photocatalytic integration, the in situ Mӧssbauer spectrum of Co0.5Fe0.5In2S4 (Fig. 5d) was conducted in different experimental condition. The isomer shift (IS=-0.08) and quadrupole splitting (QS = 0.54) with light irradiation in water solution under Ar environment is obviously different from that in the dark (IS=-0.06 and QS = 0.61). This can be understood that the unpaired valence electrons in the π-symmetry (e2g) Fe 3d-orbitals is filled by residual H* group induced by HER, which make the spin magnetic moment become smaller (Fig. 5e), leading to an influence on the Zeeman splitting. When a certain amount of nitrogen is introduced, the isomer shift and quadrupole splitting are furthered changed as IS=-0.06 and QS = 0.60. This is due to the fact that the intermediate onto the Fe and Co site is replaced by *NH2 group and their spin configurations are changed subsequently due to the orbital hybridization of Fe-N and Co-N bonds (Fig. 5e). The changes in Mӧssbauer spectra successfully confirm the photocatalytic integration between HER and NRR.