Next Article in Journal
One-Pot Synthesis for Doped Amorphous Carbon-Based Compounds: Influence of ZnO Dopant on the Charge Transfer Efficiency
Previous Article in Journal
Tunable Work Functions in Plasmonic Metals
Previous Article in Special Issue
A Review on Pre-, In-Process, and Post-Synthetic Strategies to Break the Surface Area Barrier in g-C3N4 for Energy Conversion and Environmental Remediation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bio-Inspired Photocatalytic Nitrogen Fixation: From Nitrogenase Mimicry to Advanced Artificial Systems

1
College of Biological and Chemical Engineering, Qilu Institute of Technology, Jinan 250200, China
2
Advanced Technology Research Institute, Beijing Institute of Technology, Jinan 250300, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2025, 15(19), 1485; https://doi.org/10.3390/nano15191485
Submission received: 30 July 2025 / Revised: 23 August 2025 / Accepted: 28 August 2025 / Published: 29 September 2025

Abstract

Photocatalytic nitrogen fixation under ambient conditions offers a sustainable alternative to the energy-intensive Haber–Bosch process, yet remains limited by the inertness of N≡N bonds and sluggish multi-electron/proton transfer kinetics. Nature’s nitrogenase enzymes, featuring the FeMo cofactor and ATP-driven electron cascades, inspire a new generation of artificial systems capable of mimicking their catalytic precision and selectivity. This review systematically summarizes recent advances in bio-inspired photocatalytic nitrogen reduction, focusing on six key strategies derived from enzymatic mechanisms: Fe–Mo–S active site reconstruction, hierarchical electron relay pathways, ATP-mimicking energy modules, defect-induced microenvironments, interfacial charge modulation, and spatial confinement engineering. While notable progress has been made in enhancing activity and selectivity, challenges remain in dynamic regulation, mechanistic elucidation, and system-level integration. Future efforts should prioritize operando characterization, adaptive interface design, and device-compatible catalyst platforms. By abstracting nature’s catalytic logic into synthetic architectures, biomimetic photocatalysis holds great promise for scalable, green ammonia production aligned with global decarbonization goals.

Graphical Abstract

1. Introduction: Green Ammonia and the Biomimetic Revolution

Ammonia (NH3) is a cornerstone of the global chemical industry, primarily utilized in fertilizers, but also increasingly seen as a potential hydrogen carrier and carbon-free energy vector in the transition toward net-zero emissions [1,2,3]. However, the conventional Haber–Bosch process, responsible for more than 150 million tons of annual NH3 production, consumes ~2% of the world’s total energy supply and emits over 1.5% of global CO2 emissions due to its dependence on fossil fuels and high-temperature/high-pressure conditions [4,5,6]. This contradiction between ammonia’s industrial importance and its unsustainable synthesis route has catalyzed intensive exploration of alternative green production methods [7].
Among the various routes proposed, photocatalytic nitrogen reduction reaction (NRR) under ambient conditions has emerged as one of the most attractive solutions, enabling solar-driven conversion of atmospheric N2 into NH3 without external hydrogen or harsh conditions [8,9]. Despite growing interest, most artificial photocatalysts still suffer from a low NH3 yield, which stem from the inertness of the N≡N triple bond (941 kJ mol−1) and the complex multi-electron/proton transfer process [10,11].
Nature, however, offers an elegant and efficient solution: nitrogenase enzymes, particularly the molybdenum-dependent nitrogenase featuring the FeMo cofactor (FeMoco) [12,13,14]. These enzymes catalyze N2 reduction at ambient conditions with high selectivity and turnover frequency, facilitated by a cascade of multi-metallic clusters (P-cluster, M-cluster) and ATP-driven electron/proton transfer pathways [15,16,17,18,19,20]. Nitrogenase-based systems serve as blueprints for a new generation of artificial catalysts, inspiring the development of bio-inspired photocatalysts that mimic the structural, electronic, and functional features of the natural enzyme [21,22,23].
In recent years, a surge of research has explored strategies to emulate nitrogenase catalysis using engineered nanomaterials. These include constructing metal–sulfur clusters analogous to FeMoco [24,25], designing electron relay pathways akin to P→M clusters [26,27], and incorporating defect structures to mimic flexible reaction pockets [28,29]. More advanced systems even integrate quantum dots or molecular photosensitizers to mimic the ATP-driven energy supply in nitrogenase catalysis [30]. Collectively, these systems incorporate “bio-logic” features such as spatially separated charge centers, multi-site synergy, and adaptive proton/electron relay—pushing photocatalytic NRR systems toward biomimetic performance [31,32].
This review aims to provide a comprehensive and systematic overview of the current advances in bio-inspired photocatalytic nitrogen fixation, with a particular focus on the design strategies, material innovations, and mechanistic insights that draw from nitrogenase chemistry. We first outline the biological foundation of N2 fixation, detailing the structural and functional principles of natural nitrogenases. Next, we classify and discuss key design strategies in biomimetic photocatalyst development, including FeMo active-site simulation, electron pathway engineering, artificial ATP-mimicking systems, defect regulation, and multi-site synergism. Finally, we summarize critical challenges and offer perspectives for future research that may enable scalable, sustainable, and enzyme-level efficient photocatalytic ammonia production.

2. Industrial and Biological Context of Ammonia Synthesis

2.1. Societal and Industrial Applications of Ammonia

Ammonia is one of the most important chemicals in modern society, with annual global production exceeding 180 million metric tons. Its dominant use lies in fertilizers (urea, ammonium nitrate, and ammonium sulfate), which sustain food production for more than 40% of the world’s population. Beyond agriculture, ammonia serves as a critical feedstock for diverse chemicals, including plastics, explosives, and pharmaceuticals.
In recent years, ammonia has also attracted growing attention as a carbon-free energy vector. With a high hydrogen content (17.6 wt%) and established global infrastructure for storage and transport, ammonia can act as a hydrogen carrier or be directly used in solid oxide fuel cells and combustion engines. Pilot-scale projects are already investigating ammonia co-firing in power plants and its role in green hydrogen economies.
These multifaceted applications reinforce the importance of developing sustainable N2 fixation technologies. Photocatalytic ammonia synthesis, if realized at scale, could provide not only decentralized fertilizer production but also a renewable pathway toward future energy systems.

2.2. From Haber–Bosch Process to Nitrogenase Catalysis

The Haber–Bosch process, established in the early 20th century, remains the cornerstone of industrial ammonia production. The overall reaction follows:
N2(g) + 3H2(g) ⇌ 2NH3(g), ΔH = −92.4 kJ mol−1
Despite its apparent simplicity, this reaction is thermodynamically and kinetically challenging. Activation of the N≡N triple bond requires high temperature (400–500 °C) and high pressure (150–300 bar), typically over Fe-based catalysts promoted with K2O, CaO, or Al2O3. The H2 feedstock is primarily derived from steam reforming of methane (CH4 + H2O → CO + 3H2), which contributes substantially to CO2 emissions. As a result, Haber–Bosch consumes nearly 1–2% of global energy supply and contributes more than 300 Mt CO2 annually.
A schematic of the process (Figure 1) illustrates the main stages: (i) CH4 reforming and H2 generation, (ii) N2 separation from air, (iii) high-temperature and high-pressure catalytic synthesis, and (iv) continuous recycling of unreacted gases. While this process sustains global agriculture, its extreme energy demands and environmental footprint highlight the urgent need for sustainable alternatives such as biomimetic photocatalytic N2 fixation.
The industrial Haber–Bosch process remains unmatched in terms of ammonia yield, with global plants typically producing 1500–2000 tons NH3 per day at an energy cost of 8–12 MWh per ton NH3, depending on feedstock and plant efficiency (Table 1). The reliance on methane-derived hydrogen also makes it carbon intensive, emitting 1.6–2.0 tons CO2 per ton NH3 produced. While the cost per ton is relatively low (≈300–600 USD), the environmental externalities are significant.
In contrast, current bio-inspired and photocatalytic nitrogen fixation systems achieve yields in the order of μmol–mmol NH3 g−1 h−1, far from industrial scales. However, they operate under ambient temperature and pressure without fossil-derived hydrogen input, offering the potential for decentralized, renewable-powered ammonia synthesis with drastically reduced marginal costs once efficient catalysts and device integration are realized. Recent techno-economic analyses suggest that if photocatalytic systems achieve ≥10 mmol g−1 h−1 rates with solar-to-ammonia efficiencies of 1–2%, the levelized cost of ammonia could become competitive with Haber–Bosch, particularly in regions with abundant solar resources.

3. Natural Nitrogenase Systems: Structural and Mechanistic Inspirations

Biological nitrogen fixation is accomplished by nitrogenase enzymes—nature’s only known catalysts capable of converting inert N2 into NH3 under ambient conditions. Among the three known nitrogenase isoforms (Mo, V, and Fe-only), the Mo-dependent nitrogenase is the most well-studied and catalytically efficient [15,16]. It consists of two key protein components: the dinitrogenase reductase (Fe protein) and the dinitrogenase (MoFe protein). The Fe protein mediates ATP-dependent electron delivery, while the MoFe protein houses the active site FeMo cofactor (FeMoco), a complex [MoFe7S9C] cluster responsible for N2 activation and reduction [17,18].

3.1. Structural Insights: FeMoco and P-Cluster Architectures

The FeMoco cofactor is composed of a Mo atom coordinated by three sulfur atoms and a homocitrate ligand, linked to a [Fe7S9C] cluster containing a central interstitial carbon atom. This unique cofactor is embedded within the MoFe protein, forming a precisely oriented reaction pocket that allows for N2 binding, H+ delivery, and e transfer in a concerted manner [19]. Prior to FeMoco, the electron relay occurs through the P-cluster, an [Fe8S7] double-cubane structure that bridges the Fe protein and the FeMoco site [20]. The P-cluster undergoes conformational and redox changes during the electron transfer process, functioning as a molecular switch or gate [33].
High-resolution crystallographic studies (1.0–1.3 Å) have revealed the spatial arrangement and flexibility of these clusters, shedding light on how enzyme-controlled confinement and geometry affect catalysis [18,34]. Such hierarchical structures have motivated efforts to emulate multi-metallic clusters, coordinated ligand environments, and flexible cavity-based activation sites in artificial systems.

3.2. Mechanistic Models: From Lowe–Thorneley to New Paradigms

The classic Lowe–Thorneley (LT) model describes nitrogenase catalysis as a stepwise, proton-coupled electron transfer (PCET) process, in which the FeMo-cofactor (FeMoco) cycles through a series of reduced states (E0E8) to accumulate the 8e/8H+ equivalents required for N2 fixation [21]. During the early stages (E0E4), electrons and protons delivered from the Fe protein via the P-cluster ([8Fe–7S]) are stored as bridging hydrides within FeMoco. Once sufficient reducing equivalents are accumulated, N2 binding occurs, most likely at an Fe6 site adjacent to the interstitial carbide (Figure 2).
The activation of N2 is exceptionally challenging due to the high bond dissociation energy of the N≡N triple bond (941 kJ mol−1) and the need for precise temporal control of multi-electron/proton delivery. Stepwise PCET weakens the N≡N bond through formation of key intermediates (e.g., N2H, N2H2, N2H4), eventually yielding two NH3 molecules. This process is frequently accompanied by obligatory H2 evolution, reflecting the delicate balance between desired N2 reduction and competing side reactions [35].
Recent experimental and computational insights have expanded the LT framework, suggesting alternative pathways involving early N2 binding, side-on coordination geometries, and the critical role of the interstitial carbide in tuning Fe electronic states [22]. Together, these advances emphasize not only the unique structural features of FeMoco but also the pivotal electron-relay function of the P-cluster, offering mechanistic lessons for the rational design of biomimetic and artificial NRR catalysts.
When benchmarked against the two natural and industrial references, artificial photocatalytic systems remain at an early stage. Nitrogenase catalyzes N2 reduction under ambient conditions with remarkable selectivity but requires ATP hydrolysis, limiting its turnover frequency. In contrast, Haber–Bosch achieves industrially relevant yields but at the cost of extreme temperature–pressure input and substantial carbon emissions. Artificial systems currently achieve μmol-scale production rates and modest quantum efficiencies under ambient conditions, highlighting both their unique promise and critical gap. Thus, bridging the enzymatic selectivity of nitrogenase with the scalability of Haber–Bosch remains the central challenge for next-generation catalysts (Table 2).

3.3. Role of ATP and Electron Delivery Pathways

A hallmark feature of biological nitrogen fixation is the ATP-coupled electron transfer from the Fe protein to the MoFe protein. Each electron transfer requires the hydrolysis of two ATP molecules, serving not only as an energy source but also to trigger structural rearrangement and docking of the protein units [16,17]. This energy input facilitates efficient e injection into the P-cluster and subsequently into the FeMoco site [19].
In artificial systems, mimicking this complex ATP-driven cascade remains a grand challenge. Nonetheless, the concept of spatially controlled electron flow and gated transfer between cofactor analogs has led to the use of photosensitizers, semiconductor–metal junctions, and redox mediators in bio-inspired catalyst design.

3.4. Design Implications for Artificial Systems

The architecture and function of nitrogenase enzymes provide several pivotal insights for designing biomimetic photocatalysts: (i) active center mimicry: replicating the geometry and electronic structure of FeMoco using multi-metallic clusters (e.g., Fe–Mo–S motifs); (ii) redox relay networks: constructing P-cluster-like units or conductive linkers for sequential electron transfer; (iii) cavity and ligand effects: engineering confined or defect-rich sites that facilitate N2 adsorption and PCET; (iv) energy coupling strategies: integrating ATP-analog pathways using light, voltage, or redox molecules [36,37,38,39]. These design elements are increasingly reflected in advanced catalyst architectures discussed in Section 3 and Section 4 [40,41].

4. Bio-Inspired Catalyst Design Strategies for Photocatalytic NRR

4.1. Biomimetic Construction of FeMo-like Active Centers

In biological nitrogen fixation, the FeMo cofactor (FeMoco), a highly complex [MoFe7S9C] cluster, serves as the catalytic epicenter of nitrogenase enzymes [17,18]. It integrates heteronuclear Fe–Mo active sites, sulfur-bridged coordination, and confined electronic environments, enabling stepwise proton-coupled multi-electron N≡N bond cleavage under ambient conditions. Mimicking such a multifunctional unit in artificial photocatalysts is a grand challenge, yet crucial for designing enzyme-like activity and selectivity [23,24,42].

4.1.1. Bionic Fe–Mo Architectures on Layered Semiconductors

To mimic the MoFe synergistic centers in FeMoco, researchers have constructed heteroatomic Fe/Mo frameworks embedded in low-dimensional semiconductors. Li et al. designed a bio-inspired catalyst by doping Fe atoms into two-dimensional MoTe2 nanosheets, forming an in-built “MoFe-cofactor” that emulates the redox-active nature of FeMoco (Figure 3a) [43]. The Fe dopants modulate the MoTe2 electronic structure via mixed-valence redox couples (Fe3+/Fe2+ and Mo6+/Mo4+), enhancing photogenerated electron–hole separation and prolonging the carrier lifetime by 15-fold. This significantly boosts photocatalytic N2 reduction activity—exhibiting an 11-fold enhancement over pristine MoTe2. The system represents a functional MoFe analog in a 2D inorganic matrix, retaining photo-electrochemical synergy through lattice-level redox engineering [25,42].
In a similar approach, Li et al. reported a bionic “FeV-cofactor” catalyst by doping Fe into BiVO4 and decorating it with few-layer black phosphorus (BP) nanosheets (Figure 3b) [44]. The Fe centers act as biomimetic N2 activation sites, while BP enhances interfacial charge transfer and broadens visible-light absorption. Density functional theory (DFT) analysis revealed that Fe doping facilitates N2 adsorption and activation, while V4+/V5+ and Fe2+/Fe3+ couples promote charge redistribution. The resulting FeBiVO4–0.05BP achieved a high NH3 yield of 337.9 μmol g−1 h−1, ranking among the best oxide-based photocatalysts. This study underlines the importance of cofactor-inspired multi-metal synergy and redox-coupled band modulation.

4.1.2. Atomically Precise Fe–Mo–S Motifs on Sulfide Matrices

To mimic the precise Fe–S–Mo bonding geometry in FeMoco, researchers have introduced Fe or Mo atoms into sulfur-rich matrices with atomic-level control. Wang et al. synthesized Mo(IV)-doped FeS2 nanocrystals, where Mo atoms replace Fe sites to form [Mo–S–Fe] analogs of the enzymatic cofactor (Figure 3c) [45]. The Mo dopants not only suppress competing hydrogen evolution but also enrich the active site density and promote N2 adsorption. The resulting electrocatalyst achieved a Faradaic efficiency of 14.41% at −0.2 V vs. RHE, highlighting the effectiveness of Mo-centered electron delocalization within an Fe–S lattice. This reflects the essential role of heteronuclear metal synergy and sulfur coordination in facilitating N≡N bond weakening.
In another approach, Zheng et al. engineered a catalyst where Fe atoms are anchored onto monolayer MoS2, creating [Fe–S2–Mo] motifs on a molecularly thin support (Figure 3d) [46]. This configuration emulates the structural pattern of FeMoco, with Fe atoms occupying Mo sites and forming trigonal prismatic arrangements with sulfur. Photocatalytic N2 reduction under visible light was significantly enhanced at elevated temperatures, reaching 0.24% solar-to-ammonia conversion efficiency—the highest reported among inorganic systems. The results demonstrate that spatially confined metal–sulfur coordination environments can replicate enzymatic charge localization and orbital overlap essential for N2 activation [27,46,47,48].

4.1.3. Synthetic Cubic Fe–Mo–S Clusters Emulating FeMoco Geometry

Going beyond surface modification, Ohki et al. constructed discrete [Mo3S4Fe] cubes, a molecular-level mimic of the FeMoco polyhedral core (Figure 3e) [48]. These synthetic metal–sulfur clusters feature Fe sites embedded within sulfur-rich Mo-based polyhedra, capable of capturing and reducing N2 via silylation. Under mild reductive conditions (Na/Me3SiCl), the system catalyzed N2 conversion to N(SiMe3)3, demonstrating true molecular-level N2 activation by synthetic Fe–S clusters. This work provides unprecedented evidence that Fe in sulfur-rich synthetic cubes can perform key steps in nitrogenase-like catalysis, validating the relevance of FeMoco-inspired geometry in small-molecule chemistry.

4.1.4. Design Insights and Structure–Function Correlation

Collectively, these biomimetic studies highlight several converging design principles: (i) Fe/Mo bimetallic centers replicate the redox and orbital cooperativity in FeMoco; (ii) S-rich coordination frameworks emulate the cysteine-anchored environment of enzymes; (iii) atomic-scale dispersion and facet tuning allow for charge delocalization and intermediate stabilization; and (iv) cluster-like 3D architecture introduces spatial control over multi-site electron transfer [49,50]. These strategies open up new avenues for rational design of enzyme-like photocatalysts, with enhanced N2 adsorption, reduced activation energy, and favorable multi-electron pathways under solar illumination [51,52,53,54].

4.2. Mimicking Electron Transfer Pathways of Nitrogenase

In nitrogenase catalysis, electrons are shuttled from reduced ferredoxin or flavodoxin to the Fe protein, passed through the P-cluster (Fe8S7), and finally delivered to the active FeMoco site (MoFe7S9C). This spatially organized, redox-gated, and multistage electron transfer chain is central to the enzyme’s ability to overcome the thermodynamic barrier of N≡N triple bond cleavage. Mimicking this hierarchical electron delivery pathway—without the use of ATP—is a major goal for constructing efficient artificial photocatalysts.

4.2.1. Single-Site Heterojunctions for Stepwise Electron Relay

To reproduce such directional redox gradients, Wang et al. developed a metal–organic framework (MOF)-based photocatalyst by anchoring Ru single atoms onto a Mo-MOF with a [Mo8O26(im)2]4− structure, forming Ruδ+–O3–Mo3 single-site heterojunctions (Figure 4a) [55]. These Ru–Mo ensembles mimic the P-cluster to FeMoco electron pathway by enabling sequential intraframework electron transfer. Energy band tuning was achieved by co-engineering Mo5+ defects and Ruδ+ active sites, enhancing visible-light absorption up to 700 nm and elevating the Fermi level, which reduces the energy barrier for *NNH formation—identified via DFT as the rate-limiting step. The system showed significantly higher activity than monodispersed analogs, highlighting the importance of site-specific redox heterojunctions in constructing enzyme-inspired relays.

4.2.2. Electron-Confinement Nanozymes for Redox Cycling

While biological nitrogenase relies on precise protein–protein interfaces for electron transfer, Zhang et al. introduced a bio-inspired nanozyme architecture emulating this behavior through metal oxide–oxide coupling (Figure 4b) [56]. Their CeOx/Mn3O4 system mimics symbiotic electron transfer between rhizobia and plant cells. CeOx nanoclusters (Ce3+/Ce4+ redox couple) are immobilized on root-like Mn3O4 supports (Mn2+/Mn3+ pair), enabling dynamic interfacial charge transfer analogous to enzyme–enzyme electron relay. The system exhibits catalase-like and superoxide dismutase-like activity, suggesting it can serve as a modular redox relay system in more complex N2 reduction cascades when adapted into multi-step photocatalytic platforms.

4.2.3. Dual-Site Photocathodes for Synergistic H• and e Delivery

To simulate the proton/electron coupling mechanism of nitrogenase, Zhou et al. constructed a dual-site photocathode based on atomic Mo1 sites embedded in hydrogen-substituted graphdiyne (HsGDY) anchored onto Cu2O nanowires (Figure 4c) [57]. The HsGDY scaffold, with its extended π-conjugation and high proton conductivity, acts analogously to nitrogenase’s proton relay channels, enabling directed delivery of hydrogen radicals (H•). Meanwhile, the Mo1 centers offer preferential N2 adsorption and activation. Operando spectroscopy and DFT studies confirm that the Mo1–HsGDY interface provides a low-energy dual-pathway for N≡N hydrogenation. The device achieves a remarkable NH3 yield of 78.9 μg·cm2·h1 and Faradaic efficiency of 38.9% under 10-sun illumination—among the highest in photocathodic systems—showing that enzyme-inspired spatial cooperativity at the molecular level can be realized in scalable artificial systems.

4.2.4. Hybrid Redox Scaffolds with Tandem Charge Flow

Extending this concept to hybrid materials, Tang et al. synthesized a composite catalyst comprising UiO-66(Zr–Fe) MOF integrated with Fe-dispersed MXene (Ti3C2) (Figure 4d) [58]. This architecture replicates the spatially resolved redox triad of nitrogenase: (i) UiO-66(Zr–Fe) provides photosensitization and initial electron generation; (ii) MXene nanosheets offer a conductive highway for charge migration; and (iii) atomically dispersed Fe sites act as terminal acceptors and N2 activators.
The resulting bi-phase gas–vapor–solid system achieves NH3 formation rates of up to 133.2 μmol·g−1·h−1 under full-spectrum light. The performance arises from effective photothermal coupling, multi-electron extraction, and minimized recombination loss, collectively emulating nitrogenase’s modular yet synchronized electron transfer system.

4.2.5. Mechanistic Insights and Design Implications

Across these systems, several unifying design strategies emerge (Table 3).
These bio-inspired designs reinforce the importance of electron delivery geometry, energy level alignment, and dynamic valence control in constructing next-generation photocatalysts [59,60,61]. However, unlike nitrogenase, artificial systems lack real-time conformational flexibility, precise gating, and feedback control. Bridging these remaining gaps is the challenge addressed in the next section on energy-driven systems (Section 3.3) [62,63,64].

4.3. Artificial Energy Supply Systems: Toward ATP-Free Bio-Inspired Nitrogen Reduction

In the natural nitrogenase system, each electron transfer from Fe protein to the MoFe cofactor relies on the hydrolysis of two ATP molecules, which not only provides the energy required for bond rearrangement but also coordinates the spatiotemporal electron delivery essential for multi-step N2 fixation. Emulating this sophisticated biochemical regulation in artificial systems remains a fundamental challenge. Recent studies have explored ATP-free approaches by constructing multi-component light-driven platforms, quantum–biomolecule hybrids, and membrane-inspired architectures that collectively replicate the core functionalities of ATP-fueled electron transfer and energy modulation [16,17,19].

4.3.1. Coupled Photocatalytic Hydrogen Generation and Nitrogen Reduction

A promising approach involves dual-catalyst systems that spatially decouple light-induced electron generation and nitrogen hydrogenation. In a representative study by Zhang et al., a bio-inspired photocatalytic platform was developed using H4SiW12O40 (SiW12) as a redox-active photocatalyst and Pt/C nanoparticles as hydrogenation sites (Figure 5a) [65]. Here, renewable alcohols serve as the proton and electron source. Upon visible-light irradiation, SiW12 oxidizes alcohols and stores photogenerated electrons within its polyoxometalate anion structure. These stored electrons are subsequently transferred to Pt sites to reduce adsorbed N2 via an associative–distal mechanism.
Importantly, the system exhibits a markedly low energy barrier for N–H bond formation and is capable of producing [NNHn]+ and [NHn]+ intermediates under ambient conditions. By breaking the classical scaling relations through controlled redox mediation and electron flow separation, this approach simulates ATP-triggered FeMoCo electron influx with an artificial but robust electron reservoir.

4.3.2. Semiconductor–Nitrogenase Biohybrids: Replacing ATP with Light-Driven Electron Transfer

A second approach directly couples photosensitizers with nitrogenase enzymes, offering a light-driven replacement for ATP hydrolysis. In a landmark work by Brown et al., CdS nanocrystals were used to photosensitize the MoFe protein of nitrogenase, forming a hybrid complex capable of catalyzing N2 reduction under illumination (Figure 5c,d) [68]. Here, light replaces ATP as the energy source, initiating site-specific electron transfer from CdS to MoFe clusters. The system achieved a turnover rate of 75 min−1, which is approximately 63% of that in ATP-coupled enzymatic conditions.
This hybrid model reproduces key features of the natural enzyme, including inhibition by CO and acetylene, confirming that the same mechanistic pathway is engaged. The work represents one of the most faithful mimics of biological nitrogen fixation using non-ATP pathways [67,69].

4.3.3. Purple Membrane–Ceria Hybrids: Membrane-Inspired Electron Guidance

Beyond enzymatic strategies, Jang et al. reported a photosynthetic biohybrid system (PBS) that integrates free-standing ceria nanoparticles with the purple membrane (PM) of Halobacterium salinarum (Figure 5b) [66]. Even after separation from living cells, the PM retained the ability to direct photogenerated charge carriers unidirectionally. Electrostatic self-assembly between ceria and PM resulted in close interfacial contact, facilitating efficient N2-to-NH3 conversion under ambient pressure and temperature without requiring any toxic or precious metals.
In this construct, the Ce3+/Ce4+ redox pair serves as an intrinsic electron shuttle, while PM functions as a photonic membrane to regulate electron orientation. This platform mimics the spatial energy regulation provided by ATP binding and release cycles, enabling selective and efficient abiotic NRR in aqueous environments.

4.3.4. Design Outlook

These three systems highlight complementary ATP-free routes toward biomimetic NRR catalysis. Key insights include the following: (i) Electron storage–release systems (e.g., polyoxometalates) can serve as artificial equivalents of ATP hydrolysis cycles. (ii) Semiconductor–enzyme hybrids allow light to directly fuel biological catalysts with preserved selectivity. (iii) Biohybrid membranes can emulate directional energy transfer and coupling through internal redox pairs [70,71].
Collectively, they embody a shift from energy-intensive thermal systems to self-sustained, ambient-condition NRR architectures that reflect the elegance of biological catalysis while advancing toward scalable application [72,73,74,75].

5. Interface Engineering and Defect Control in Biomimetic Photocatalysts

5.1. Bio-Inspired Structural Motifs: Bridging Atomic Architectures and Photocatalytic Activity

Nature’s nitrogen fixation mechanism, primarily driven by the FeMo cofactor in nitrogenase, has provided powerful inspiration for the design of heterogeneous photocatalysts [76]. This section summarizes recent progress in mimicking structural features of biological systems—such as coordination environments, vacancy architectures, and multi-metal active centers—to enhance N2 activation and hydrogenation efficiency under mild conditions [42].

5.1.1. Zr–O Clusters in MOFs: Mimicking MoFe Binding Domains

Guo et al. designed a UiO-66(SH)2-based MOF whose Zr6O6 cluster plays a role analogous to the MoFe cluster in nitrogenase, with a controlled dehydration step enabling μ-N–Zr interactions (Figure 6a) [77]. The stepwise weakening of the N≡N bond was attributed to photo-induced electrons facilitating successive protonation inside the [Zr6O6] cage. In situ EXAFS and TGA confirmed structural dynamics post-dehydration, and 15N2 isotopic labeling validated ammonia as the exclusive product. This study highlights how metal–oxygen polyhedra can be designed to mimic the reaction cavity and binding specificity of metalloenzymes.

5.1.2. Light-Switchable Oxygen Vacancies in Bismuth-Based Nanotubes

Inspired by enzyme flexibility and redox plasticity, Wang et al. created ultrathin Bi5O7Br nanotubes (5 nm) with controllable surface oxygen vacancies (OVs), which act as dynamic electron reservoirs for activating N2 (Figure 6b) [78]. These light-switchable OVs enabled a high ammonia yield of 1.38 mmol h−1 g−1 and a quantum efficiency exceeding 2.3% under 420 nm illumination in pure water. The unique tubular morphology and large surface area promote N2 adsorption, while visible-light-induced OVs simulate the redox plasticity of nitrogenase. This work offers an elegant example of mimicking the electron-rich environments of enzyme pockets using nanostructured defect sites.

5.1.3. Mo-Doped Hematite: Electrochemical Analog of FeMo Cofactor

Niu et al. synthesized Mo-doped α-Fe2O3 porous nanospheres exhibiting Fe–O–Mo subunits that emulate the electronic synergy of the FeMo cofactor (Figure 6c) [79]. Mo doping modified the morphology and enriched defect density, enhancing N2 chemisorption and electron conductivity. Electrochemical NRR experiments revealed a Faradaic efficiency of 11.2% and NH3 yield of 21.3 ± 1.1 μg h−1 mgcat−1 under ambient conditions. DFT simulations showed that Mo doping elongated the N≡N bond (1.132 Å) and favored an associative distal pathway. This result suggests that atomic-level heteroatom engineering is essential to translate enzymatic motifs into scalable electrocatalytic platforms [18,33].

5.1.4. Polysulfide-Based Bimetallic Centers: Synergistic Fe–V Activation

Cai et al. constructed a biomimetic photocatalyst by incorporating a vanadium-doped iron polysulfide active center onto nitrogen-deficient g-C3N5, mimicking the multi-metal synergism of nitrogenase (Figure 6d) [80]. While Fe in the polysulfide component preferentially adsorbed dinitrogen, V facilitated water activation and proton transfer. This dual-function system showed significantly boosted N2 fixation activity (1916.5 μmol h−1 g−1), 6.81 times higher than the pristine g-C3N5. The spatial separation of N2 activation and protonation zones emulates enzymatic pathways and reduces charge recombination [81].

5.1.5. Design Implications

These examples illustrate how defect engineering can serve biomimetic functions: (i) electronic tuning of the band edge via Ov/Nv/Sv; (ii) dual-site chemisorption for cooperative N2 binding and hydrogenation; (iii) formation of enzyme-like cavities that regulate intermediate stabilization; and (iv) charge localization and multi-electron storage, mimicking enzymatic redox buffering [82,83,84].
As shown in Figure 4, the above defect-engineered materials introduce localized asymmetry, orbital perturbation, and catalytic specificity that resemble nitrogenase cofactor microenvironments. In Section 4.2, we explore how interfacial regulation further enhances these effects by creating heterojunctions and cooperative interfaces [85,86].

5.2. Interfacial Engineering: Constructing Functional Junctions for Catalytic Enhancement

In biological nitrogen fixation, interprotein interfaces play an indispensable role in ensuring spatiotemporal control of electron delivery and protonation. Inspired by this [16,17,19], artificial photocatalytic systems have focused on constructing heterointerfaces to manipulate charge separation, regulate reaction energetics, and enable dual-function catalysis [44].

5.2.1. Biological Membrane–Nanoparticle Hybrid: Interface-Guided Redox Modulation

To extend interfacial engineering into hybrid bioinorganic systems, Jang et al. developed a photosynthetic biohybrid system (PBS) by integrating free-standing CeO2 nanoparticles into purple membranes (PMs) derived from Halobacterium salinarum (Figure 7a) [66]. Unlike the previous structural focus in Section 3.3, the emphasis here is on interface-directed charge modulation [30,87].
The electrostatic interaction between PM and CeO2 established unidirectional electron transfer channels under visible light, enabling efficient photoreduction of N2 to NH3 at room temperature and atmospheric pressure—without requiring noble metals or enzyme reconstitution. This interfacial assembly mirrors the protein–cofactor coupling in natural nitrogenase, offering a stable and tunable microenvironment that enables redox versatility while preserving catalytic integrity [18,33].

5.2.2. In-Lattice Heterojunctions: Atomic-Level Carrier Highways

Chen et al. reported an “in-lattice heterojunction” photocatalyst by embedding Fe–N4-anchored carbon (Fe–N–C) layers into the cavity of defective TiO2 (D-TiO2), forming a confined Fe–N–C@D-TiO2 structure with atomic continuity along the Ti–C–N–Fe axis (Figure 7b) [81]. This arrangement established radial charge carrier migration pathways with minimized recombination and enhanced electronic communication [25].
More importantly, the tailored heterojunction promoted N2 adsorption and activation at Fe–N4 centers, leading to an ammonia yield of 88 μmol g−1 h−1—significantly surpassing that of non-cavity Fe–N–C catalysts. The lattice-confined configuration mimics the biological Fe protein–MoFe protein complex in terms of confined spatial orientation and charge steering via coordinated atomic channels [25,30].

5.2.3. Tandem Interfaces: Hot Electron-Driven Multistage Catalysis

Yang et al. introduced a tandem photocatalytic system consisting of plasmonic gold nanocrystals (Au NPs) anchored on ultrathin TiO2 nanosheets with abundant oxygen vacancies (OVs) (Figure 7c,d) [88]. The system exhibited dual interfacial functionality: the TiO2 vacancies chemisorbed and activated N2 molecules, while hot electrons generated by localized surface plasmon resonance (LSPR) of Au NPs reduced the activated intermediates [58].
This “working-in-tandem” configuration led to a high ammonia production rate with an apparent quantum efficiency of 0.82% at 550 nm. Moreover, optimized combinations of Au nanospheres and nanorods enhanced broadband light utilization, improving the N2 fixation rate by 66.2%. The spatial division of excitation and activation resembles enzymatic compartmentalization within nitrogenase’s multi-domain architecture [29,69].

5.2.4. Facet-Dependent Electronic Structures and N2 Adsorption

Different crystal facets of a photocatalyst expose specific atomic terminations, which define surface energy, electron density, and adsorbate binding behavior [29]. Li et al. constructed {001}/{010} co-exposed BiVO4 nanosheets and found that the {010} facet promotes charge accumulation, while the {001} facet preferentially adsorbs N2, leading to spatially resolved *N2 reduction and *H2O oxidation [44]. This mirrors how FeMoco’s geometry creates selective substrate tunnels to separate proton and N2 channels [89,90]. Similarly, Liu et al. synthesized TiO2 nanocrystals with dominant {101} facets, which display enhanced *N2 chemisorption and lower work function compared to {001} facets [83]. This reflects how enzymes use facet-like pocket polarity to achieve differential substrate interactions [91].

5.2.5. Summary of Biomimetic Interfacial Strategies

Across these representative systems, several biomimetic features emerge: (i) atomic-scale charge migration pathways, emulating protein electron tunnels; (ii) dual-function tandem interfaces, enabling sequential excitation and activation; and (iii) biohybrid membranes, integrating organic–inorganic interfaces for redox control [29,69]. Together, these strategies reflect the essence of biological nitrogen fixation interfaces—spatial precision, redox flexibility, and selective electron–proton coupling—offering insights for future design of efficient N2 fixation catalysts under ambient conditions [92,93].

6. Conclusions and Outlook

Biomimetic photocatalytic nitrogen fixation represents a compelling strategy to bridge the performance chasm between natural enzymatic nitrogenase systems and artificial N2-to-NH3 conversion technologies. Through rational mimicry of structural, electronic, and functional features of nitrogenase, recent advances have demonstrated how catalyst design can transcend traditional semiconductor limitations and approach biological efficiency and selectivity under ambient conditions [94].
This review systematically dissects six core bio-inspired design strategies (Figure 2, Figure 3, Figure 4, Figure 5 and Figure 6): (i) active site mimicry (Fe–S, Mo–Fe–C, Fe-only cofactor models) reconstructs catalytic cores akin to FeMoco [43,44,45,65,77]; (ii) electron transfer pathways simulate P-cluster → FeMoco relay cascades via π-conductive networks and redox modules [55,58,59,60]; (iii) artificial energy supply systems bypass ATP via plasmonic, redox cycling, and light-driven donor–acceptor modules [60,65,66,68]; (iv) defect engineering introduces asymmetry, charge localization, and dual-site activation resembling enzymatic pocket dynamics [76,77,79,80,81]; (v) interfacial regulation mimics enzymatic docking, redox cooperativity, and molecular orientation via semiconducting–molecular hybrids [44,58,78,88]; and (vi) facet and spatial confinement strategies replicate enzymatic channel gating, microenvironmental steering, and intermediate stabilization [44,48,66,81].
Despite these achievements, several fundamental challenges remain: (i) Inadequate dynamic modulation: Unlike nitrogenase, which orchestrates redox steps via ATP-induced conformational gating [18,20,33], most artificial systems lack temporal regulation mechanisms. Static structures fail to match the dynamic demands of proton–electron coupling required for multi-step N2 reduction [25,69]. (ii) Ambiguity in reaction mechanism: The exact activation route—associative distal, alternating, or enzymatic hybrid pathways—often remains speculative. Although DFT modeling is widely used [60,79,81], experimental evidence such as isotope tracing ([15N], [1H] NMR), in-situ XAFS/FTIR, and operando spectroscopy is still scarce [26,69,87]. (iii) Poor integration in system-level devices: Most reported photocatalytic systems operate in suspension or half-cell modes. Translation into scalable, stable device platforms (e.g., flow reactors, PEC systems) is minimal. Integration with solar-to-ammonia systems or bio-hybrid platforms (quantum dot–enzyme conjugates [68]) remains in early stages.
To push the frontier of biomimetic photocatalytic NRR, future research may focus on the following: (i) Dynamic Interface Engineering: Develop stimuli-responligands, pH-sensitive redox gates) that mimic ATP-regulated Fe-protein docking [17,19]. Temporal control of electron flux may drastically improve N2RR turnover efficiency. (ii) Mechanistic Decoding by Multimodal Characterization: Combine DFT with operando Raman, Mössbauer, and isotopic tracing to distinguish real intermediates (e.g., *NNH, *NH2) and validate simulated reaction coordinates. Synchronizing theory and spectroscopy will clarify pathway ambiguity [26,29,87]. (iii) Biohybrid and Device Integration: Construct MOF–nanozyme–semiconductor hybrids or quantum dot–cofactor mimetics that maintain enzyme-like function within scalable architectures. Integrate into photochemical or PEC systems with bias-free N2RR, approaching real-world application. (vi) AI-Guided Catalyst Design: Utilize machine learning with descriptors such as Bader charge, orbital occupancy, and vacancy formation energy to guide cofactor-level screening and facet–defect–interface synergies.
In summary, although each strategy provides unique advantages, hybrid approaches that combine the structural fidelity of biomimetic catalysts, the scalability of semiconductor vacancy engineering, and the directional charge transfer of biohybrids appear most promising. Particularly, atomic-level site engineering and tandem plasmonic–semiconductor systems stand out for bridging efficiency and scalability. The convergence of these approaches with advanced in situ characterization, AI-driven screening, and device-level integration may yield the next generation of nitrogenase-inspired catalysts capable of sustainable ammonia synthesis under ambient conditions.

Author Contributions

Conceptualization, L.C. and Z.G.; methodology, W.X., K.Z., J.H. and H.F.; validation, W.X. and K.Z.; formal analysis, W.X., K.Z., J.H. and H.F.; investigation, W.X., K.Z., J.H., H.F. and M.G.; resources, H.-Z.H., Y.L.P., H.M. and X.G.; data curation, W.X., K.Z. and J.H.; writing—original draft preparation, W.X. and K.Z.; writing—review and editing, L.C., S.H. and Z.G.; visualization, K.Z., J.H. and H.F.,; supervision, L.C., S.H. and Z.G.; project administration, L.C. and S.H.; funding acquisition, L.C., S.H., Y.L.P. and Z.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Shandong Provincial Natural Science Foundation (ZR2024QB122, ZR2024QE448), the National Natural Science Foundation of China (82202240), the Taishan Scholar Young Talent Program (tsqn202211249), and the Research Program of Qilu Institute of Technology (No. QIT23TP011).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Jia, H.P.; Quadrelli, E.A. Mechanistic aspects of dinitrogen cleavage and hydrogenation to produce ammonia in catalysis and organometallic chemistry: Relevance of metal hydride bonds and dihydrogen. Chem. Soc. Rev. 2014, 43, 547–564. [Google Scholar] [CrossRef]
  2. Li, J.; Chen, R.; Wang, J.; Zhou, Y.; Yang, G.; Dong, F. Subnanometric alkaline-earth oxide clusters for sustainable nitrate to ammonia photosynthesis. Nat. Commun. 2022, 13, 1098. [Google Scholar] [CrossRef] [PubMed]
  3. Zhao, Y.; Wu, F.; Miao, Y.; Zhou, C.; Xu, N.; Shi, R.; Wu, L.Z.; Tang, J.; Zhang, T. Revealing ammonia quantification minefield in photo/electrocatalysis. Angew. Chem. Int. Ed. 2021, 60, 21728. [Google Scholar] [CrossRef]
  4. Erisman, J.W.; Sutton, M.A.; Galloway, J.; Klimont, Z.; Winiwarter, W. How a century of ammonia synthesis changed the world. Nat. Geosci. 2008, 1, 636–639. [Google Scholar] [CrossRef]
  5. Li, L.; Jiang, Y.F.; Zhang, T.; Cai, H.; Lin, B.; Lin, X.; Zheng, Y.; Zheng, L.; Wang, X.; Xu, C.Q.; et al. Size sensitivity of supported Ru catalysts for ammonia synthesis: From nanoparticles to subnanometric clusters and atomic clusters. Chem 2022, 8, 749–768. [Google Scholar] [CrossRef]
  6. Wang, Q.; Lei, Y.; Wang, D.; Li, Y. Defect engineering in earth-abundant electrocatalysts for CO2 and N2 reduction. Energ. Environ. Sci. 2019, 12, 1730–1750. [Google Scholar] [CrossRef]
  7. Liu, T.T.; Zhai, D.D.; Guan, B.T.; Shi, Z.J. Nitrogen fixation and transformation with main group elements. Chem. Soc. Rev. 2022, 51, 3846–3861. [Google Scholar] [CrossRef]
  8. Zhang, L.; Hou, S.; Wang, T.; Liu, S.; Gao, X.; Wang, C.; Wang, G. Recent advances in application of graphitic carbon nitride-based catalysts for photocatalytic nitrogen fixation. Small 2022, 18, 2202252. [Google Scholar] [CrossRef]
  9. Liu, H.; Timoshenko, J.; Bai, L.; Li, Q.; Rüscher, M.; Sun, C.; Cuenya, B.R.; Luo, J. Low-coordination rhodium catalysts for an efficient electrochemical nitrate reduction to ammonia. ACS Catal. 2023, 13, 1513–1521. [Google Scholar] [CrossRef]
  10. Chen, J.G.; Crooks, R.M.; Seefeldt, L.C.; Bren, K.L.; Bullock, R.M.; Darensbourg, M.Y.; Holland, P.L.; Hoffman, B.; Janik, M.J.; Jones, A.K.; et al. Beyond fossil fuel-driven nitrogen transformations. Science 2018, 360, eaar6611. [Google Scholar] [CrossRef] [PubMed]
  11. Schrauzer, G.N.; Guth, T.D. Photolysis of water and photoreduction of nitrogen on titanium dioxide. J. Am. Chem. Soc. 1977, 99, 7189–7193. [Google Scholar] [CrossRef]
  12. de Barros, H.R.; Garcia, I.; Kuttner, C.; Zeballos, N.; Camargo, P.H.C.; de Torresi, S.I.C.; López-Gallego, F.; Liz-Marzán, L.M. Mechanistic insights into the light-driven catalysis of an immobilized lipase on plasmonic nanomaterials. ACS Catal. 2020, 11, 414–423. [Google Scholar] [CrossRef]
  13. Shang, S.; Xiong, W.; Yang, C.; Johannessen, B.; Liu, R.; Hsu, H.Y.; Gu, Q.; Leung, M.K.H.; Shang, J. Atomically dispersed iron metal site in a porphyrin-based metal-organic framework for photocatalytic nitrogen fixation. ACS Nano 2021, 15, 9670–9678. [Google Scholar] [CrossRef]
  14. Utomo, W.P.; Leung, M.K.H.; Yin, Z.; Wu, H.; Ng, Y.H. Advancement of bismuth-based materials for electrocatalytic and photo (electro) catalytic ammonia synthesis. Adv. Funct. Mater. 2022, 32, 2106713. [Google Scholar] [CrossRef]
  15. Huang, Y.; Zhu, Y.; Chen, S.; Xie, X.; Wu, Z.; Zhang, N. Schottky junctions with bi cocatalyst for taming aqueous phase N2 reduction toward enhanced solar ammonia production. Adv. Sci. 2021, 8, 2003626. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, Y.; Zhao, J.; Yang, D.; Wang, B.; Zhou, Y.; Wang, J.; Chen, H.; Mei, T.; Ye, S.; Qu, J. A thiolate- bridged FeIVFeIV μ-nitrido complex and its hydrogenation reactivity toward ammonia formation. Nat. Chem. 2022, 14, 46–52. [Google Scholar] [CrossRef]
  17. Ryu, M.H.; Zhang, J.; Toth, T.; Khokhani, D.; Geddes, B.A.; Mus, F.; Garcia-Costas, A.; Peters, J.W.; Poole, P.S.; AnéJ, M.; et al. Control of nitrogen fixation in bacteria that associate with cereals. Nat. Microbiol. 2020, 5, 314–330. [Google Scholar] [CrossRef]
  18. Einsle, O.; Rees, D.C. Structural enzymology of nitrogenase enzymes. Chem. Rev. 2020, 120, 4969–5004. [Google Scholar] [CrossRef]
  19. Jasniewski, A.J.; Lee, C.C.; Ribbe, M.W.; Hu, Y. Reactivity, mechanism, and assembly of the alternative nitrogenases. Chem. Rev. 2020, 120, 5107–5157. [Google Scholar] [CrossRef]
  20. Lowe, D.J.; Thorneley, R.N. The mechanism of Klebsiella pneumoniae nitrogenase action. Pre- steady-state kinetics of H2 formation. Biochem. J. 1984, 224, 877–886. [Google Scholar] [CrossRef]
  21. Buscagan, T.M.; Rees, D.C. Rethinking the nitrogenase mechanism: Activating the active site. Joule 2019, 3, 2662–2678. [Google Scholar] [CrossRef] [PubMed]
  22. Scott, D.J.; May, H.D.; Newton, W.E.; Brigle, K.E.; Dean, D.R. Role for the nitrogenase MoFe protein α-subunit in FeMo-cofactor binding and catalysis. Nature 1990, 343, 188–190. [Google Scholar] [CrossRef] [PubMed]
  23. Chica, B.; Ruzicka, J.; Kallas, H.; Mulder, D.W.; Brown, K.A.; Peters, J.W.; Seefeldt, L.C.; Dukovic, G.; King, P.W. Defining intermediates of nitrogenase MoFe protein during N2 reduction under photochemical electron delivery from CdS quantum dots. J. Am. Chem. Soc. 2020, 142, 14324–14330. [Google Scholar] [CrossRef]
  24. Brown, K.A.; Ruzicka, J.; Kallas, H.; Chica, B.; Mulder, D.W.; Peters, J.W.; Seefeldt, L.C.; Dukovic, G.; King, P.W. Excitation-rate determines product stoichiometry in photochemical ammonia production by CdS quantum dot-nitrogenase MoFe protein complexes. ACS Catal. 2020, 10, 11147–11152. [Google Scholar] [CrossRef]
  25. Ruzicka, J.L.; Pellows, L.M.; Kallas, H.; Shulenberger, K.E.; Zadvornyy, O.A.; Chica, B.; Brown, K.A.; Peters, J.W.; King, P.W.; Seefeldt, L.C.; et al. The kinetics of electron transfer from CdS nanorods to the MoFe protein of nitrogenase. Indian. J. Chem. A 2022, 126, 8425–8435. [Google Scholar] [CrossRef]
  26. Zander, J.; Timm, J.; Weiss, M.; Marschall, R. Light-induced ammonia generation over defective carbon nitride modified with pyrite. Adv. Energy Mater. 2022, 12, 2202403. [Google Scholar] [CrossRef]
  27. Zhang, Y.; Wang, Q.; Yang, S.; Wang, H.; Rao, D.; Chen, T.; Wang, G.; Lu, J.; Zhu, J.; Wei, S.; et al. Tuning the interaction between ruthenium single atoms and the second coordination sphere for efficient nitrogen photofixation. Adv. Funct. Mater. 2022, 32, 2112452. [Google Scholar] [CrossRef]
  28. Han, T.; Cao, X.; Sun, K.; Peng, Q.; Ye, C.; Huang, A.; Cheong, W.C.; Chen, Z.; Lin, R.; Zhao, D.; et al. Anion-exchange-mediated internal electric field for boosting photogenerated carrier separation and utilization. Nat. Commun. 2021, 12, 4952. [Google Scholar] [CrossRef]
  29. Wang, G.; Huang, R.; Zhang, J.; Mao, J.; Wang, D.; Li, Y. Synergistic modulation of the separation of photo-generated carriers via engineering of dual atomic sites for promoting photocatalytic performance. Adv. Mater. 2021, 33, 2105904. [Google Scholar] [CrossRef]
  30. Zhang, F.; Liu, Y.; Liu, J.; Wu, H.; Jiang, Y.; Zhou, H.; Guo, H.; Zheng, K.; Yartsev, A.; Zhou, Y. Engineering bimetallic CuAu to suppress hydrogen evolution and enhance charge transfer for improved photoelectrochemical nitrogen reduction. Appl. Catal. B-Environ. 2025, 26, 125739. [Google Scholar] [CrossRef]
  31. Chang, F.; Gao, W.; Guo, J.; Chen, P. Emerging materials and methods toward ammonia-based energy storage and conversion. Adv. Mater. 2021, 33, 2005721. [Google Scholar] [CrossRef] [PubMed]
  32. Fang, Z.; Wu, P.; Qian, Y.; Yu, G. Gel-derived amorphous bismuth-nickel alloy promotes electrocatalytic nitrogen fixation via optimizing nitrogen adsorption and activation. Angew. Chem. Int. Ed. 2021, 60, 4275–4281. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, X.; You, J.; Xue, Y.; Ren, J.; Zhang, K.; Fu, B.; Xue, Q.; Tian, J.; Zhang, H. Structural regulation of three-dimensional bismuth vanadate nanochannels for excellent visible light photocatalytic nitrogen fixation. Appl. Catal. B-Environ. 2025, 363, 124817. [Google Scholar] [CrossRef]
  34. Lee, C.C.; Kang, W.; Jasniewski, A.J.; Stiebritz, M.T.; Tanifuji, K.; Ribbe, M.W.; Hu, Y. Evidence of substrate binding and product release via belt-sulfur mobilization of the nitrogenase cofactor. Nat. Catal. 2022, 5, 443–454. [Google Scholar] [CrossRef]
  35. Stappen, C.V.; Decamps, L.; Cutsail III, G.E.; Bjornsson, R.; Henthorn, J.T.; Birrell, J.A.; DeBeer, S. The spectroscopy of nitrogenases. Chem. Rev. 2020, 120, 5005–5081. [Google Scholar] [CrossRef]
  36. Huang, P.W.; Hatzell, M.C. Prospects and good experimental practices for photocatalytic ammonia synthesis. Nat. Commun. 2022, 13, 7908. [Google Scholar] [CrossRef]
  37. Guan, Y.; Wen, H.; Cui, K.; Wang, Q.; Gao, W.; Cai, Y.; Cheng, Z.; Pei, Q.; Li, Z.; Cao, H.; et al. Light-driven ammonia synthesis under mild conditions using lithium hydride. Nat. Chem. 2024, 16, 373. [Google Scholar] [CrossRef]
  38. Yin, H.; Chen, Z.; Peng, Y.; Xiong, S.; Li, Y.; Yamashita, H.; Li, J. Dual active centers bridged by oxygen vacancies of ruthenium single-atom hybrids supported on molybdenum oxide for photocatalytic ammonia synthesis. Angew. Chem. Int. Ed. 2022, 61, e202114242. [Google Scholar] [CrossRef]
  39. Mao, C.; Wang, J.; Zou, Y.; Shi, Y.; Viasus, C.J.; Loh, J.Y.Y.; Xia, M.; Ji, S.; Li, M.; Shang, H.; et al. Photochemical acceleration of ammonia production by Pt1-Ptn-TiN reduction and N2 activation. J. Am. Chem. Soc. 2023, 145, 13134. [Google Scholar] [CrossRef]
  40. Shi, Y.; Zhao, Z.; Yang, D.; Tan, J.; Xin, X.; Liu, Y.; Jiang, Z. Engineering photocatalytic ammonia synthesis. Chem. Soc. Rev. 2023, 52, 6938. [Google Scholar] [CrossRef] [PubMed]
  41. He, T.; Zhao, Z.; Liu, R.; Liu, X.; Ni, B.; Wei, Y.; Wu, Y.; Yuan, W.; Peng, H.; Jiang, Z.; et al. Porphyrin-based covalent organic frameworks anchoring Au single atoms for photocatalytic nitrogen fixation. J. Am. Chem. Soc. 2023, 145, 6057. [Google Scholar] [CrossRef]
  42. Ge, X.; Zheng, X.; Zhou, T.; Tian, L.-J.; Wang, W.; Chen, J.-J.; Wang, X. Engineering spin state of atomic iron centers for high-performance photocatalytic nitrogen fixation. Angew. Chem. Int. Ed. 2025. Online ahead of print. [Google Scholar] [CrossRef]
  43. Li, H.; Gu, S.; Sun, Z.; Guo, F.; Xie, Y.; Tao, B.; He, X.; Zhang, W.; Chang, H. The in-built bionic “MoFe-cofactor” in Fe-doped two-dimensional MoTe2 nanosheets for boostingthe photocatalytic nitrogen reductionperformance. J. Mater. Chem. A 2020, 8, 13038–13048. [Google Scholar] [CrossRef]
  44. Li, H.; Jian, S.; Tao, B.; Xu, G.; Liu, B.; Gu, S.; Wang, G.; Chang, H. “FeV–cofactor”–inspired bionic Fe–doped BiVO4 photocatalyst decorated with few layer 2D black phosphorus for efficient nitrogen reduction. Inorg. Chem. Front. 2023, 10, 5004–5015. [Google Scholar] [CrossRef]
  45. Wang, H.; Wang, J.; Zhang, R.; Cheng, C.; Qiu, K.; Yang, Y.; Mao, J.; Liu, H.; Du, M.; Dong, C.; et al. Bionic design of a Mo(lV)-doped FeS, catalyst for electroreduction of dinitrogen to ammonia. ACS Catal. 2020, 10, 4914–4921. [Google Scholar] [CrossRef]
  46. Zheng, J.; Lu, L.; Lebedev, K.; Wu, S.; Zhao, P.; Pherson, I.J.; Wu, T.; Kato, R.; Li, Y.; Ho, P.; et al. Fe on molecular-layer MoS2 as inorganic Fe-S2-Mo motifs for light-driven nitrogen fixation to ammonia at elevated temperatures. Chem Catal. 2021, 1, 162–182. [Google Scholar] [CrossRef]
  47. Dong, G.; Qiu, P.; Chen, Q.; Huang, C.; Chen, F.; Liu, X.; Li, Z.; Wang, Y.; Zhao, Y. K+-intercalated carbon nitride with electron storage property for high-efficiency visible light driven nitrogen fixation. Chem. Eng. J. 2022, 433, 133573. [Google Scholar] [CrossRef]
  48. Ohki, Y.; Munakata, K.; Matsuoka, Y.; Hara, R.; Kachi, M.; Uchida, K.; Tada, M.; Cramer, R.E.; Sameera, W.M.C.; Takayama, T.; et al. Nitrogen reduction by the Fe sites of synthetic [Mo3S4Fe] cubes. Nature 2022, 607, 86–90. [Google Scholar] [CrossRef]
  49. Kong, Y.; Li, X.; Santiago, A.R.P.; He, T. Nonmetal atom doping induced orbital shifts and charge modulation at the edge of two-dimensional boron carbonitride leading to enhanced photocatalytic nitrogen reduction. J. Am. Chem. Soc. 2024, 146, 5987. [Google Scholar] [CrossRef]
  50. Yuan, J.; Feng, W.; Zhang, Y.; Xiao, J.; Zhang, X.; Wu, Y.; Ni, W.; Huang, H.; Dai, W. Unraveling synergistic effect of defects and piezoelectric field in breakthrough piezo-photocatalytic N2 reduction. Adv. Mater. 2024, 36, 2303845. [Google Scholar] [CrossRef]
  51. Wu, R.; Gao, S.; Jones, C.; Sun, M.; Guo, M.; Tai, R.; Chen, S.; Wang, Q. Bi/BSO heterojunctions via vacancy engineering for efficient photocatalytic nitrogen fixation. Adv. Funct. Mater. 2024, 34, 2314051. [Google Scholar] [CrossRef]
  52. Xiong, Y.; Li, B.; Gu, Y.; Yan, T.; Ni, Z.; Li, S.; Zuo, J.L.; Ma, J.; Jin, Z. Photocatalytic nitrogen fixation under an ambient atmosphere using a porous coordination polymer with bridging dinitrogen anions. Nat. Chem. 2023, 15, 286. [Google Scholar] [CrossRef] [PubMed]
  53. Li, H.; Zhang, J.; Deng, X.; Wang, Y.; Meng, G.; Liu, R.; Huang, J.; Tu, M.; Xu, C.; Peng, Y.; et al. Structure and defect engineering synergistically boost high solar-to-chemical conversion efficiency of cerium oxide/Au hollow nanomushrooms for nitrogen photofixation. Angew. Chem. Int. Ed. 2024, 63, e202316384. [Google Scholar] [CrossRef]
  54. Yuan, Y.J.; Lu, N.; Bao, L.; Tang, R.; Zhang, F.G.; Guan, J.; Wang, H.D.; Liu, Q.Y.; Cheng, Q.; Yu, Z.T.; et al. SiP nanosheets: A metal-free two-dimensional photocatalyst for visible-light photocatalytic H2 production and nitrogen fixation. ACS Nano 2022, 16, 12174. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, S.; Zhang, J.; Zong, M.; Xu, J.; Wang, D.; Bu, X. Energy level engineering: Ru single atom anchored on Mo-MOF with a [Mo8O26(im)2]4- structure acts as a biomimetic photocatalyst. ACS Catal. 2022, 12, 7960–7974. [Google Scholar] [CrossRef]
  56. Zhang, J.; Wang, Z.H.; Lin, X.; Gao, X.; Wang, Q.; Huang, R.; Ruan, Y.; Xu, H.; Tian, L.; Ling, C.; et al. Mn-Ce symbiosis: Nanozymes with multiple active sites Facilitate scavenging of reactive oxygen species (ROS) basedon electron transfer and confinement anchoring. Angew. Chem. Int. Ed. 2025, 64, e202416686. [Google Scholar] [CrossRef]
  57. Zhou, X.; Wu, W.; Xu, D.; Liu, C.; Zhang, Z. Bioinspired dual-site photocathode with atomic molybdenum and alkynyl networks for scalable solar ammonia synthesis. ACS Nano 2025, 19, 19417–19428. [Google Scholar] [CrossRef]
  58. Tang, Y.; Liu, Z.; Jia, J.; Zeng, H.; Rui, Z. Accelerating electron transfer dynamics for efficient nitrogen photoreduction on nitrogenase-mimicking metal-organic framework/Fe-dispersed MXene. Appl. Catal. B-Environ. 2024, 358, 124426. [Google Scholar] [CrossRef]
  59. Dong, X.; Cui, Z.; Shi, X.; Yan, P.; Wang, Z.; Co, A.C.; Dong, F. Insights into dynamic surface bromide sites in Bi4O5Br2 for sustainable N2 photofixation. Angew. Chem. Int. Ed. 2022, 61, e202200937. [Google Scholar] [CrossRef]
  60. Li, X.H.; Li, H.; Jiang, S.L.; Yang, L.; Li, H.Y.; Liu, Q.L.; Bai, W.; Zhang, Q.; Xiao, C.; Xie, Y. Constructing mimic-enzyme catalyst: Polyoxometalates regulating carrier dynamics of metal–organic frameworks to promote photocatalytic nitrogen fixation. ACS Catal. 2023, 13, 7189. [Google Scholar] [CrossRef]
  61. Li, X.; Dong, Q.; Wang, J.; Liang, X.; Robertson, P.K.J.; Li, F.; Guo, M.; Choi, W.; Wang, C. Molecular engineering of active Fe center in metalloporphyrin coupled with polyoxometalates for efficient photochemical nitrogen fixation: Synergistic effect of multiactive sites strengthening metal-N-N* interactions. Adv. Funct. Mater. 2025, 35, 2424128. [Google Scholar] [CrossRef]
  62. Wang, X.; Fan, G.; Guo, S.; Gao, R.; Guo, Y.; Han, C.; Gao, Y.; Zhang, J.; Gu, X.; Wu, L. Regulated dual defects of bridging organic and terminal inorganic ligands in iron-based metal-organic framework nodes for efficient photocatalytic ammonia synthesis. Angew. Chem. Int. Ed. 2024, 63, e202404258. [Google Scholar] [CrossRef]
  63. Wang, X.; Gao, R.; Fan, G.; Guo, Y.; Han, C.; Gao, Y.; Shen, A.; Wu, L.; Gu, X. Dual defects-induced iron single atoms immobilized in metal–organic framework-derived hollow BiOBr microtubes for low-barrier photocatalytic nitrogen reduction. Angew. Chem. Int. Ed. 2025, 64, e202501297. [Google Scholar] [CrossRef]
  64. Bo, Y.; Wang, H.; Lin, Y.; Yang, T.; Ye, R.; Li, Y.; Hu, C.; Du, P.; Hu, Y.; Liu, Z.; et al. Altering hydrogenation pathways in photocatalytic nitrogen fixation by tuning local electronic structure of oxygen vacancy with dopant. Angew. Chem. Int. Ed. 2021, 60, 16085. [Google Scholar] [CrossRef] [PubMed]
  65. Zhang, R.; Sun, W.; Zhang, Y.; Feng, Y.; Dai, B.; Deng, Y. Photocatalytic hydrogen production from alcohols to in-situ hydrogenation of nitrogen to ammonia by a dual-catalyst system. Appl. Catal. B-Environ. 2024, 358, 124429. [Google Scholar] [CrossRef]
  66. Jang, J.; Liu, Y.; Gosztola, D.J.; Kuzmenko, I.; Niklas, J.; Poluektov, O.G.; Lee, B.; Rozhkova, E.A. Photosynthetic biohybrid system for enhanced abiotic N2-to-NH3 conversion under ambient conditions. J. Am. Chem. Soc. 2025. Online ahead of print. [Google Scholar] [CrossRef] [PubMed]
  67. An, K.; Ren, H.; Yang, D.; Zhao, Z.; Gao, Y.; Chen, Y.; Tan, J.; Wang, W.; Jiang, Z. Nitrogenase-inspired bimetallic metal organic frameworks for visible-light-driven nitrogen fixation. Appl. Catal. B-Environ. 2021, 292, 120167. [Google Scholar] [CrossRef]
  68. Brown, K.A.; Harris, D.F.; Wilker, M.B.; Rasmussen, A.; Khadka, N.; Hamby, H.; Keable, S.; Dukovic, G.; Peters, J.W.; Seefeldt, L.C.; et al. Light-driven dinitrogen reduction catalyzed by a CdS:nitrogenase MoFe protein biohybrid. Science 2016, 352, 448–450. [Google Scholar] [CrossRef]
  69. Kang, W.; Lee, C.C.; Jasniewski, A.J.; Ribbe, M.W.; Hu, Y. Structural evidence for a dynamic metallocofactor during N2 reduction by Mo-nitrogenase. Science 2020, 368, 1381–1385. [Google Scholar] [CrossRef]
  70. Dong, G.; Huang, X.; Bi, Y. Anchoring black phosphorus quantum dots on Fe-Doped W18O49 nanowires for efficient photocatalytic nitrogen fixation. Angew. Chem. Int. Ed. 2022, 61, e202204271. [Google Scholar] [CrossRef]
  71. Di, J.; Chen, C.; Wu, Y.; Zhao, Y.; Zhu, C.; Zhang, Y.; Wang, C.; Chen, H.; Xiong, J.; Xu, M.; et al. Polarized Cu–Bi site pairs for non-covalent to covalent interaction tuning toward N2 photoreduction. Adv.; Mater. 2022, 34, 2204959. [Google Scholar] [CrossRef]
  72. Sun, Y.; Ji, H.; Sun, Y.; Zhang, G.; Zhou, H.; Cao, S.; Liu, S.; Zhang, L.; Li, W.; Zhu, X.; et al. Synergistic effect of oxygen vacancy and high porosity of nano MIL-125(Ti) for enhanced photocatalytic nitrogen fixation. Angew. Chem. Int. Ed. 2024, 63, e202316973. [Google Scholar] [CrossRef] [PubMed]
  73. Li, J.; Zhang, C.; Bao, T.; Xi, Y.; Yuan, L.; Zou, Y.; Bi, Y.; Liu, C.; Yu, C. Dual near-infrared-response S-scheme heterojunction with asymmetric adsorption sites for enhanced nitrogen photoreduction. Adv. Mater. 2025, 37, 2416210. [Google Scholar] [CrossRef] [PubMed]
  74. Wang, X.; Zhao, Y.; Wu, X.; Zhang, B.; Tian, J.; Wong, W.Y.; Zhang, F. Achieving photocatalytic overall nitrogen fixation via an enzymatic pathway on a distorted CoP4 configuration. Angew. Chem. Int. Ed. 2025, 137, e202420327. [Google Scholar] [CrossRef]
  75. Wang, J.; Ran, G.; Gao, J.; Li, D.; Waterhouse, G.I.N.; Shi, R.; Zhang, W.; Tang, J.; Wu, L.Z.; Zhao, Y.; et al. Solar-driven conversion of nitrogen and water to solid fertilizer in an outdoor 1 m2 panel reactor. Adv. Mater. 2025, 37, 2420199. [Google Scholar] [CrossRef]
  76. Sun, B.; Lu, S.; Qian, Y.; Zhang, X.; Tian, J. Recent progress in research and design concepts for the characterization, testing, and photocatalysts for nitrogen reduction reaction. Carbon Energy 2023, 5, e305. [Google Scholar] [CrossRef]
  77. Guo, B.; Cheng, X.; Tang, Y.; Guo, W.; Deng, S.; Wu, L.; Fu, X. Dehydrated UiO-66(SH)2: The Zr-O cluster and its photocatalytic role mimicking the biological nitrogen fixation. Angew. Chem. Int. Ed. 2022, 61, e202117244. [Google Scholar] [CrossRef]
  78. Wang, S.; Hai, X.; Ding, X.; Chang, K.; Xiang, Y.; Meng, X.; Yang, Z.; Chen, H.; Ye, J. Light-switchable oxygen vacancies in ultrafine Bi5O7Br nanotubes for boosting solar-driven nitrogen fixation in pure water. Adv. Mater. 2017, 29, 1701774. [Google Scholar] [CrossRef]
  79. Niu, Z.; Jiao, L.; Zhang, T.; Zhao, X.; Wang, X.; Tan, Z.; Liu, L.; Chen, S.; Song, X. Boosting electrocatalytic ammonia synthesis of bio-inspired porous Mo-doped hematite via nitrogen Activation. ACS Appl. Mater. Interfaces 2022, 14, 55559–55567. [Google Scholar] [CrossRef]
  80. Cai, W.; Li, K.; Wen, J.; Zhang, Z.; Zhong, Q.; Qu, H. The construction of a vanadium-doped polysulfide iron bimetallic active center for enhanced photocatalytic nitrogen fixation. Catal. Sci. Technol. 2023, 13, 5567–5575. [Google Scholar] [CrossRef]
  81. Chen, T.Y.; Ying, Y.R.; Wu, J.; Liu, X.H.; Huang, H. Fe-N4-anchored carbon layer patched TiO2 cavities to construct an “in-lattice heterojunction” for enhanced photocatalytic nitrogen reduction reactions. Angew. Chem. 2025. Online ahead of print. [Google Scholar] [CrossRef]
  82. Guan, R.; Cheng, X.; Chen, Y.; Wu, Z.; Zhao, Z.; Shang, Q.; Sun, Y.; Sun, Z. Wettability control of defective TiO2 with alkyl acid for highly efficient photocatalytic ammonia synthesis. Nano Res. 2023, 16, 10770. [Google Scholar] [CrossRef]
  83. Liu, C.; Hao, D.; Ye, J.; Ye, S.; Zhou, F.; Xie, H.; Qin, G.; Xu, J.; Liu, J.; Li, S.; et al. Knowledge-driven design and lab-based evaluation of B-doped TiO2 photocatalysts for ammonia synthesis. Adv. Energy Mater. 2023, 13, 2204126. [Google Scholar] [CrossRef]
  84. Ren, G.; Zhao, J.; Zhao, Z.; Li, Z.; Wang, L.; Zhang, Z.; Li, C.; Meng, X. Defects-induced single-atom anchoring on metal–organic frameworks for high-efficiency photocatalytic nitrogen reduction. Angew. Chem. Int. Ed. 2023, 63, e202314408. [Google Scholar] [CrossRef] [PubMed]
  85. Yuan, L.; Tang, C.; Du, P.; Li, J.; Zhang, C.; Xi, Y.; Bi, Y.; Bao, T.; Du, A.; Liu, C.; et al. Nanoporous heterojunction photocatalysts with engineered interfacial sites for efficient photocatalytic nitrogen fixation. Angew. Chem. Int. Ed. 2024, 63, e202412340. [Google Scholar] [CrossRef] [PubMed]
  86. Zhao, Z.; Nian, Y.; Shi, J.; Xin, X.; Huang, X.; Shi, Y.; Tan, J.; Zhang, Y.; Han, Y.; Yang, D.; et al. Engineering reaction pathway to harmonize the competition between NRR and HER for efficient photocatalytic ammonia synthesis. ACS Catal. 2024, 14, 11626. [Google Scholar] [CrossRef]
  87. Li, X.; Hai, G.; Liu, J.; Zhao, F.; Peng, Z.; Liu, H.; Leung, M.K.H.; Wang, H. Bio-inspired NiCoP/CoMoP/Co(Mo3Se4)4@C/NF multi-heterojunction nanoflowers: Effective catalytic nitrogen reduction by driving electron transfer. Appl. Catal. B-Environ. 2022, 314, 121531. [Google Scholar] [CrossRef]
  88. Yang, J.; Guo, Y.; Jiang, R.; Qin, F.; Zhang, H.; Lu, W.; Wang, J.; Yu, J.C. High-efficiency “working-in-tandem” nitrogen photofixation achieved by assembling plasmonic gold nanocrystals on ultrathin titania nanosheets. J. Am. Chem. Soc. 2018, 140, 8497–8508. [Google Scholar] [CrossRef]
  89. Gayatri, S.; Sabiha, S.; Kulamani, P. Constructing a novel surfactant-free MoS2 nanosheet modified MgIn2S4 marigold microflower: An efficient visible-light driven 2D-2D p-n heterojunction photocatalyst toward HER and pH regulated NRR. ACS. Sustain. Chem. Eng. 2020, 12, 4848–4862. [Google Scholar]
  90. Gao, N.; Yang, H.; Dong, D.; Dou, D.; Liu, Y.; Zhou, W.; Gao, F.; Nan, C.; Liang, Z.; Yang, D. Bi2S3 quantum dots in situ grown on MoS2 nanoflowers: An efficient electron-rich interface for photoelectrochemical N2 reduction. J. Colloid Interf. Sci. 2022, 611, 294–305. [Google Scholar] [CrossRef]
  91. Gong, X.; Chong, B.; Xia, M.; Li, H.; Ou, H.; Yang, G. Fluorinated phosphonium ionic liquid boosts the N2-adsorbing ability of TiO2 for efficient photocatalytic NH3 synthesis. Catal. Sci. Technol. 2024, 14, 343. [Google Scholar] [CrossRef]
  92. Rafieerad, A.; Amiri, A.; Yan, W.; Eshghi, H.; Dhingra, S. Conversion of 2D MXene to multi-low-dimensional GerMXene superlattice heterostructure. Adv. Funct. Mater. 2022, 32, 2108495. [Google Scholar] [CrossRef] [PubMed]
  93. Wang, H.; Seemakurthi, R.R.; Chen, G.F.; Strauss, V.; Savateev, O.; Hai, G.; Ding, L.; López, N.; Wang, H.; Antonietti, M. Laser-induced nitrogen fixation. Nat. Commun. 2023, 14, 5668. [Google Scholar] [CrossRef] [PubMed]
  94. Liu, S.; Wang, M.; Ji, H.; Zhang, L.; Ni, J.; Li, N.; Qian, T.; Yan, C.; Lu, J. Solvent-in-gas system for promoted photocatalytic ammonia synthesis on porous framework materials. Adv. Mater. 2023, 35, 2211730. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Simplified schematic of the Haber–Bosch process.
Figure 1. Simplified schematic of the Haber–Bosch process.
Nanomaterials 15 01485 g001
Figure 2. Schematic illustration of N2 reduction and electron transfer within the nitrogenase complex.
Figure 2. Schematic illustration of N2 reduction and electron transfer within the nitrogenase complex.
Nanomaterials 15 01485 g002
Figure 3. Representative strategies for active-site mimicry in photocatalytic nitrogen fixation: structural and electronic emulation of the FeMo cofactor. Reproduced with permission. (a) Proposed photocatalytic N2 reduction pathway over a Fe–MoTe2 heterostructure under visible light. The Fe–Mo dual-site interface replicates the binuclear arrangement of FeMoco, enabling synergistic adsorption and activation of N2 via charge redistribution. Reproduced with permission [43]. Copyright © 2020 Royal Society of Chemistry. (b) Schematic representation of the FeBiVO4-0.05BP photocatalyst. The Fe active site simulates the iron center of nitrogenase, while BiVO4 provides a visible-light-absorbing matrix. The dual-functional architecture facilitates stepwise N2 protonation and suppresses competing HER via proton channel redirection. Adapted from ref. Reproduced with permission [44]. Copyright © 2023 Royal Society of Chemistry. (c) Molecular and structural design of Mo(IV)-doped FeS2. Upper panel: structural analogy to the FeMo cofactor of nitrogenase with [Fe–S2–Mo] motifs; lower panel: schematic of FeS2 with Mo substitution, which induces Fe–Mo–S coordination resembling FeMoco subunits and enhances N2 adsorption capacity. Reproduced with permission [45]. Copyright © 2020 American Chemical Society. (d) Atomic-scale imaging and model of Fe single-atom anchored on sMoS2. The Fe1 center occupies a Mo atop site in a trigonal prismatic environment, forming a [Fe–S2–Mo] unit analogous to enzymatic cofactor geometry. The system exhibits strong photothermal synergy in NRR, with photogenerated excitons extended by thermally activated carrier lifetimes. Reproduced with permission [46]. Copyright © 2021 Cell Press. (e) Schematic of K+-intercalated crystalline carbon nitride (KCCN) containing nitrogen vacancies (NVs) for visible-light-driven nitrogen photofixation. NVs chemisorb and activate N2, while intercalated K+ enables robust electron storage within the carbon nitride framework [47]. Copyright © 2022 Elsevier.
Figure 3. Representative strategies for active-site mimicry in photocatalytic nitrogen fixation: structural and electronic emulation of the FeMo cofactor. Reproduced with permission. (a) Proposed photocatalytic N2 reduction pathway over a Fe–MoTe2 heterostructure under visible light. The Fe–Mo dual-site interface replicates the binuclear arrangement of FeMoco, enabling synergistic adsorption and activation of N2 via charge redistribution. Reproduced with permission [43]. Copyright © 2020 Royal Society of Chemistry. (b) Schematic representation of the FeBiVO4-0.05BP photocatalyst. The Fe active site simulates the iron center of nitrogenase, while BiVO4 provides a visible-light-absorbing matrix. The dual-functional architecture facilitates stepwise N2 protonation and suppresses competing HER via proton channel redirection. Adapted from ref. Reproduced with permission [44]. Copyright © 2023 Royal Society of Chemistry. (c) Molecular and structural design of Mo(IV)-doped FeS2. Upper panel: structural analogy to the FeMo cofactor of nitrogenase with [Fe–S2–Mo] motifs; lower panel: schematic of FeS2 with Mo substitution, which induces Fe–Mo–S coordination resembling FeMoco subunits and enhances N2 adsorption capacity. Reproduced with permission [45]. Copyright © 2020 American Chemical Society. (d) Atomic-scale imaging and model of Fe single-atom anchored on sMoS2. The Fe1 center occupies a Mo atop site in a trigonal prismatic environment, forming a [Fe–S2–Mo] unit analogous to enzymatic cofactor geometry. The system exhibits strong photothermal synergy in NRR, with photogenerated excitons extended by thermally activated carrier lifetimes. Reproduced with permission [46]. Copyright © 2021 Cell Press. (e) Schematic of K+-intercalated crystalline carbon nitride (KCCN) containing nitrogen vacancies (NVs) for visible-light-driven nitrogen photofixation. NVs chemisorb and activate N2, while intercalated K+ enables robust electron storage within the carbon nitride framework [47]. Copyright © 2022 Elsevier.
Nanomaterials 15 01485 g003
Figure 4. Bio-inspired design strategies for photocatalytic nitrogen fixation materials based on nitrogenase mimics. (a) Construction of single-site heterojunctions via Ru-anchored Mo-based MOFs enables precise energy band engineering. The coupling of Ru doping and Mo5+ defect states leads to enhanced light absorption and spatial charge separation, significantly boosting N2 activation. Reproduced with permission [55]. Copyright © 2022 American Chemical Society. (b) A nanozyme system mimicking plant–rhizobia symbiosis, constructed via confined CeOx clusters anchored on Mn3O4 nanorods, achieves lattice interlacing and directional electron transfer (Mn2+→Ce4+), facilitating ROS scavenging and creating a protective redox microenvironment. Reproduced with permission [56]. Copyright © 2025 Wiley-VCH. (c) Schematic design of a nitrogenase-inspired photoelectrode based on Mo1/HsGDY@Cu2O, integrating defect-rich Mo sites with highly conjugated graphdiyne to form synergistic active centers for efficient solar-driven PEC-NRR. Reproduced with permission [57]. Copyright © 2025 American Chemical Society. (d) Rational assembly of UiO-66(Zr-Fe)/Fe-decorated MXene simulates the nitrogenase multi-site tandem pathway, with respective roles in photothermal conversion, electron transport, and N2 fixation. A high NH3 yield is achieved under solar irradiation in a gas–vapor–solid triphasic system. Reproduced with permission [58]. Copyright © 2024 Elsevier.
Figure 4. Bio-inspired design strategies for photocatalytic nitrogen fixation materials based on nitrogenase mimics. (a) Construction of single-site heterojunctions via Ru-anchored Mo-based MOFs enables precise energy band engineering. The coupling of Ru doping and Mo5+ defect states leads to enhanced light absorption and spatial charge separation, significantly boosting N2 activation. Reproduced with permission [55]. Copyright © 2022 American Chemical Society. (b) A nanozyme system mimicking plant–rhizobia symbiosis, constructed via confined CeOx clusters anchored on Mn3O4 nanorods, achieves lattice interlacing and directional electron transfer (Mn2+→Ce4+), facilitating ROS scavenging and creating a protective redox microenvironment. Reproduced with permission [56]. Copyright © 2025 Wiley-VCH. (c) Schematic design of a nitrogenase-inspired photoelectrode based on Mo1/HsGDY@Cu2O, integrating defect-rich Mo sites with highly conjugated graphdiyne to form synergistic active centers for efficient solar-driven PEC-NRR. Reproduced with permission [57]. Copyright © 2025 American Chemical Society. (d) Rational assembly of UiO-66(Zr-Fe)/Fe-decorated MXene simulates the nitrogenase multi-site tandem pathway, with respective roles in photothermal conversion, electron transport, and N2 fixation. A high NH3 yield is achieved under solar irradiation in a gas–vapor–solid triphasic system. Reproduced with permission [58]. Copyright © 2024 Elsevier.
Nanomaterials 15 01485 g004
Figure 5. Biohybrid energy supply systems inspired by nitrogenase for light-driven nitrogen fixation under mild conditions. (a) A dual-catalyst photocatalytic system integrates polyoxometalate SiW12 as a photoredox mediator and Pt/C as the nitrogen hydrogenation site, enabling alcohol-driven in situ hydrogen generation and sustainable N2-to-NH3 conversion. Solar-driven dehydrogenation of alcohols supplies protons and electrons, which are stored in reduced SiW12 and subsequently transferred to Pt/C for N2 reduction via an associative distal pathway, bypassing the need for external H2. Reproduced with permission [65]. Copyright © 2024 Elsevier. (b) A photosynthetic biohybrid system (PBS) composed of ceria nanoparticles embedded within light-responsive purple membrane (PM) fragments from halobacterium salinarum. The bacteriorhodopsin chromophores (yellow rods) embedded in PM facilitate directional charge transfer under solar illumination, enabling efficient abiotic N2 reduction on ceria surfaces under ambient conditions. Reproduced with permission [66]. Copyright © 2025 American Chemical Society. (c) A nitrogenase-inspired metal-organic framework (MOF) catalyst, U(Zr-Hf)-2SH, integrates Zr-Hf bimetallic clusters with thiol-functionalized linkers. In this architecture, Zr sites serve as nitrogen activation centers while Hf acts as an electron reservoir, enabling a ligand-to-metal-to-metal electron transfer (LMMET) mechanism. The -SH-containing organic ligands extend visible-light absorption and promote π-backbonding with N2, collectively driving a high NH3 generation rate [67]. Copyright © 2021 Elsevier.
Figure 5. Biohybrid energy supply systems inspired by nitrogenase for light-driven nitrogen fixation under mild conditions. (a) A dual-catalyst photocatalytic system integrates polyoxometalate SiW12 as a photoredox mediator and Pt/C as the nitrogen hydrogenation site, enabling alcohol-driven in situ hydrogen generation and sustainable N2-to-NH3 conversion. Solar-driven dehydrogenation of alcohols supplies protons and electrons, which are stored in reduced SiW12 and subsequently transferred to Pt/C for N2 reduction via an associative distal pathway, bypassing the need for external H2. Reproduced with permission [65]. Copyright © 2024 Elsevier. (b) A photosynthetic biohybrid system (PBS) composed of ceria nanoparticles embedded within light-responsive purple membrane (PM) fragments from halobacterium salinarum. The bacteriorhodopsin chromophores (yellow rods) embedded in PM facilitate directional charge transfer under solar illumination, enabling efficient abiotic N2 reduction on ceria surfaces under ambient conditions. Reproduced with permission [66]. Copyright © 2025 American Chemical Society. (c) A nitrogenase-inspired metal-organic framework (MOF) catalyst, U(Zr-Hf)-2SH, integrates Zr-Hf bimetallic clusters with thiol-functionalized linkers. In this architecture, Zr sites serve as nitrogen activation centers while Hf acts as an electron reservoir, enabling a ligand-to-metal-to-metal electron transfer (LMMET) mechanism. The -SH-containing organic ligands extend visible-light absorption and promote π-backbonding with N2, collectively driving a high NH3 generation rate [67]. Copyright © 2021 Elsevier.
Nanomaterials 15 01485 g005
Figure 6. Bio-inspired defect engineering strategies enabling charge redistribution, electron localization, and N≡N bond activation in photocatalytic nitrogen fixation systems. (a) Stepwise structural evolution of UiO-66, UiO-66(SH)2, and its dehydrated form illustrating defect formation and photo-induced charge transfer. SH-functionalization lowers the bandgap, while dehydration triggers linker loss and generates Zr–coordinatively unsaturated sites (CUSs). These CUSs act as artificial enzyme-like coordination nodes, enabling μ-N–Zr interactions for visible-light-driven N2 reduction. Reproduced with permission [77]. Copyright © 2022 Wiley-VCH. (b) Schematic synthesis of ultrafine Bi5O7Br nanotubes from Bi–Br–oleylamine precursors. The tubular architecture enhances light absorption, exposes more defect-rich surfaces, and facilitates directional electron migration. Oxygen vacancies (OVs) at the tube wall act as N2 adsorption centers, simulating active FeMoco cavities. Reproduced with permission [78]. Copyright © 2017 Wiley-VCH. (c) Bio-inspired Mo-doped Fe2O3 porous nanospheres for electrochemical N2 reduction. Mo atoms replace Fe sites, forming Fe–O–Mo subunits resembling nitrogenase cofactors. The doping induces a porous morphology, defect generation, and favorable orbital redistribution. DFT simulations confirm N≡N bond elongation and a reduced barrier for the associative distal pathway. Reproduced with permission [79]. Copyright © 2022 American Chemical Society. (d) Proposed charge-separation mechanism in FeVS/C3N4 hybrid under visible light. The interface promotes spatially separated photogenerated electrons and holes, with electrons migrating toward FeVS to activate adsorbed N2. Conduction band alignment and defect-assisted transfer mimic the directional tunneling in enzymatic NRR. Reproduced with permission [80]. Copyright © 2023 Royal Society of Chemistry.
Figure 6. Bio-inspired defect engineering strategies enabling charge redistribution, electron localization, and N≡N bond activation in photocatalytic nitrogen fixation systems. (a) Stepwise structural evolution of UiO-66, UiO-66(SH)2, and its dehydrated form illustrating defect formation and photo-induced charge transfer. SH-functionalization lowers the bandgap, while dehydration triggers linker loss and generates Zr–coordinatively unsaturated sites (CUSs). These CUSs act as artificial enzyme-like coordination nodes, enabling μ-N–Zr interactions for visible-light-driven N2 reduction. Reproduced with permission [77]. Copyright © 2022 Wiley-VCH. (b) Schematic synthesis of ultrafine Bi5O7Br nanotubes from Bi–Br–oleylamine precursors. The tubular architecture enhances light absorption, exposes more defect-rich surfaces, and facilitates directional electron migration. Oxygen vacancies (OVs) at the tube wall act as N2 adsorption centers, simulating active FeMoco cavities. Reproduced with permission [78]. Copyright © 2017 Wiley-VCH. (c) Bio-inspired Mo-doped Fe2O3 porous nanospheres for electrochemical N2 reduction. Mo atoms replace Fe sites, forming Fe–O–Mo subunits resembling nitrogenase cofactors. The doping induces a porous morphology, defect generation, and favorable orbital redistribution. DFT simulations confirm N≡N bond elongation and a reduced barrier for the associative distal pathway. Reproduced with permission [79]. Copyright © 2022 American Chemical Society. (d) Proposed charge-separation mechanism in FeVS/C3N4 hybrid under visible light. The interface promotes spatially separated photogenerated electrons and holes, with electrons migrating toward FeVS to activate adsorbed N2. Conduction band alignment and defect-assisted transfer mimic the directional tunneling in enzymatic NRR. Reproduced with permission [80]. Copyright © 2023 Royal Society of Chemistry.
Nanomaterials 15 01485 g006
Figure 7. Biomimetic interfacial engineering strategies for enhanced photocatalytic nitrogen fixation. (a) Schematic of a purple-membrane (PM)-coated CeO2 nanoparticle system mimicking light-driven ATP-free proton/electron coupling for N2 reduction. The archaeal PM acts as a bio-light harvester, enabling photogenerated electrons to reduce N2 on ceria surfaces via an enzyme-like multi-channel reaction pathway. Organic acids (formic acid, lactic acid, glycerol) serve as hole scavengers to maintain redox balance. Reproduced with permission [66]. Copyright © 2025 American Chemical Society. (b) Comparison of enzymatic and artificial coordination environments. The upper panel illustrates the Fe–S–Mo–C cofactor in nitrogenase with ATP-powered electron transfer for *N2 reduction, while the lower panel shows a dehydrated Zr6O6-cluster-based MOF (UiO-66-(SH)2) with μ-N–Zr binding, enabling visible-light-driven NRR via a proton-assisted multi-electron process. Reproduced with permission [81]. Copyright © 2025 Wiley-VCH. (c) Schematic of the natural nitrogenase complex comprising the Fe protein and MoFe protein. ATP hydrolysis induces conformational changes at the docking interface, enabling stepwise electron delivery from the Fe protein to the MoFe protein and finally to *N2 for reduction to NH3. Reproduced with permission [88]. Copyright © 2018 American Chemical Society. (d) A tandem-interface photocatalyst composed of Au nanocrystals anchored on ultrathin TiO2 nanosheets with abundant oxygen vacancies (OVs). Hot electrons from surface plasmon resonance (Au) and chemisorption activation at OVs synergistically weaken the N≡N bond, facilitating selective NH3 formation. Apparent quantum efficiency reaches 0.82% at 550 nm, highlighting the catalytic advantages of interface synergy and visible-light harvesting. Reproduced with permission [88]. 2018 American Chemical Society.
Figure 7. Biomimetic interfacial engineering strategies for enhanced photocatalytic nitrogen fixation. (a) Schematic of a purple-membrane (PM)-coated CeO2 nanoparticle system mimicking light-driven ATP-free proton/electron coupling for N2 reduction. The archaeal PM acts as a bio-light harvester, enabling photogenerated electrons to reduce N2 on ceria surfaces via an enzyme-like multi-channel reaction pathway. Organic acids (formic acid, lactic acid, glycerol) serve as hole scavengers to maintain redox balance. Reproduced with permission [66]. Copyright © 2025 American Chemical Society. (b) Comparison of enzymatic and artificial coordination environments. The upper panel illustrates the Fe–S–Mo–C cofactor in nitrogenase with ATP-powered electron transfer for *N2 reduction, while the lower panel shows a dehydrated Zr6O6-cluster-based MOF (UiO-66-(SH)2) with μ-N–Zr binding, enabling visible-light-driven NRR via a proton-assisted multi-electron process. Reproduced with permission [81]. Copyright © 2025 Wiley-VCH. (c) Schematic of the natural nitrogenase complex comprising the Fe protein and MoFe protein. ATP hydrolysis induces conformational changes at the docking interface, enabling stepwise electron delivery from the Fe protein to the MoFe protein and finally to *N2 for reduction to NH3. Reproduced with permission [88]. Copyright © 2018 American Chemical Society. (d) A tandem-interface photocatalyst composed of Au nanocrystals anchored on ultrathin TiO2 nanosheets with abundant oxygen vacancies (OVs). Hot electrons from surface plasmon resonance (Au) and chemisorption activation at OVs synergistically weaken the N≡N bond, facilitating selective NH3 formation. Apparent quantum efficiency reaches 0.82% at 550 nm, highlighting the catalytic advantages of interface synergy and visible-light harvesting. Reproduced with permission [88]. 2018 American Chemical Society.
Nanomaterials 15 01485 g007
Table 1. Comparison of Haber–Bosch process and bio-inspired photocatalytic nitrogen fixation in terms of yield and cost.
Table 1. Comparison of Haber–Bosch process and bio-inspired photocatalytic nitrogen fixation in terms of yield and cost.
ParameterHaber–BoschBio-Inspired Photocatalysis
Typical operating conditions400–500 °C, 150–300 barAmbient temperature and pressure
Yield scale1500–2000 tons NH3 per dayμmol–mmol NH3 g−1 h−1
Energy cost8–12 MWh per ton NH3Mainly solar input
CO2 emissions1.6–2.0 tons CO2 per ton NH3Near-zero (if powered by renewables)
Cost per ton (USD)300–600 (fossil-based H2)Currently not scalable; projected competitive if ≥1–2% solar-to-ammonia efficiency is reached
Table 2. Benchmark comparison of N2-to-NH3 systems.
Table 2. Benchmark comparison of N2-to-NH3 systems.
SystemRate/EfficiencyConditionsAdvantagesLimitations
Nitrogenase (MoFe protein)~10–20 NH3 s−1 per enzyme (~10−2 mmol g−1 h−1); Faradaic efficiency ~ 60–70%Ambient T, P; ATP-drivenOperates under mild conditions, high selectivityRequires 16 ATP per N2, slow compared to industrial process
Haber–Bosch>1000 mmol gcat−1 h−1 (industrial scale); ~10–15% energy efficiency400–500 °C, 150–300 atm, Fe/Co catalystsScalable, mature, >150 Mt NH3 annuallyEnormous energy input, CO2 emissions
Artificial photocatalysts (recent)10–500 μmol g−1 h−1; quantum efficiency up to 2%Ambient T, P, solar-drivenMild, renewable-driven, tunable selectivity3–5 orders slower than Haber–Bosch; stability challenges
Table 3. Comparative mapping of key design elements in biological nitrogenase versus artificial biomimetic photocatalysts.
Table 3. Comparative mapping of key design elements in biological nitrogenase versus artificial biomimetic photocatalysts.
Design ElementBiological CounterpartArtificial Mimic
Stepwise redox sites P-cluster → FeMocoRuδ+–Mo5+ pairs in MOFs [55]
Redox cycling via metal valence Protein-based switchingCe3+/Ce4+–Mn2+/Mn3+ relay [56]
Directional H•/e supply Proton-coupled electron relayHsGDY–Mo1 dual site [57]
Tandem charge transport Ferredoxin to FeMocoUiO-66/Fe–MXene relay [58]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xia, W.; Zhang, K.; Hou, J.; Fu, H.; Gao, M.; Huang, H.-Z.; Chen, L.; Han, S.; Pak, Y.L.; Mou, H.; et al. Bio-Inspired Photocatalytic Nitrogen Fixation: From Nitrogenase Mimicry to Advanced Artificial Systems. Nanomaterials 2025, 15, 1485. https://doi.org/10.3390/nano15191485

AMA Style

Xia W, Zhang K, Hou J, Fu H, Gao M, Huang H-Z, Chen L, Han S, Pak YL, Mou H, et al. Bio-Inspired Photocatalytic Nitrogen Fixation: From Nitrogenase Mimicry to Advanced Artificial Systems. Nanomaterials. 2025; 15(19):1485. https://doi.org/10.3390/nano15191485

Chicago/Turabian Style

Xia, Wenpin, Kaiyang Zhang, Jiewen Hou, Huaiyu Fu, Mingming Gao, Hui-Zi Huang, Liwei Chen, Suqin Han, Yen Leng Pak, Hongyu Mou, and et al. 2025. "Bio-Inspired Photocatalytic Nitrogen Fixation: From Nitrogenase Mimicry to Advanced Artificial Systems" Nanomaterials 15, no. 19: 1485. https://doi.org/10.3390/nano15191485

APA Style

Xia, W., Zhang, K., Hou, J., Fu, H., Gao, M., Huang, H.-Z., Chen, L., Han, S., Pak, Y. L., Mou, H., Gao, X., & Guo, Z. (2025). Bio-Inspired Photocatalytic Nitrogen Fixation: From Nitrogenase Mimicry to Advanced Artificial Systems. Nanomaterials, 15(19), 1485. https://doi.org/10.3390/nano15191485

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop