Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Research progress on structural regulation of nitrogen-fixing photocatalysts

Zhao Zhanfeng *a, Zhang Yue b and Gao Ningning a
aSINOPEC Research Institute of Petroleum Processing, State Key Laboratory of Petroleum Molecular & Process Engineering, Beijing, China. E-mail: zhaozhanfeng.ripp@sinopec.com
bChina Wuzhou Engineering Group Corporation Ltd, First Design and Research Institute, Beijing, China

Received 9th February 2025 , Accepted 10th April 2025

First published on 16th April 2025


Abstract

Photocatalytic nitrogen fixation is a forward-looking technology for zero-carbon nitrogen fixation, which is crucial for alleviating the energy crisis and achieving carbon neutrality. Based on research into the structural regulation of nitrogen-fixing photocatalysts, this review summarizes the latest progress and challenges in photocatalytic ammonia synthesis from three dimensions: active sites, crystal structures, and composite structures. In terms of active site construction, common types of active sites, including metal sites, non-metal sites, vacancies, and single atoms, are discussed. Their characteristics and methods for improving photocatalytic nitrogen fixation performance are analyzed. Furthermore, starting from the mechanism of nitrogen activation, a general strategy for active sites to promote the electron exchange process and thereby enhance nitrogen activation efficiency is explored. In terms of crystal structure construction, the design of nitrogen-fixing photocatalysts is described from three perspectives: crystal form, crystal facet, and morphology control. In terms of composite structure construction, this review discusses the key role of structures such as semiconductor–metal composites and semiconductor–semiconductor composites in promoting carrier separation. It is hoped that this review can provide new insights for the design and preparation of efficient nitrogen-fixing photocatalysts and inspire practical applications of photocatalytic nitrogen fixation.


image file: d5ra00953g-p1.tif

Zhao Zhanfeng

Zhao Zhanfeng, born in May 1993, obtained his PhD from Tianjin University in 2023. He is currently employed at the Sinopec Research Institute of Petroleum and Petrochemicals, where his primary research focuses on photocatalytic nitrogen fixation and the structural regulation of metal–organic framework materials.

image file: d5ra00953g-p2.tif

Zhang Yue

Zhang Yue, born in April 1996, obtained her PhD from Tianjin University in 2024. She is currently working at China Wuzhou Engineering Group Corporation Ltd, with her primary research focusing on the structural regulation of catalysts.

image file: d5ra00953g-p3.tif

Gao Ningning

Gao Ningning, born in November 1985, obtained her PhD from the Dalian Institute of Chemical Physics, Chinese Academy of Sciences in 2014. She is currently working at Sinopec Research Institute of Petroleum and Petrochemicals Co., Ltd, with her primary research focusing on photocatalysis, industrial catalysis, adsorptive separation, and membrane separation.


1. Introduction

Nitrogen fixation, the conversion of free nitrogen molecules (e.g., N2) into nitrogen compounds absorbable by humans for synthesizing essential substances such as DNA, RNA, and proteins, serves as a crucial cornerstone in maintaining ecological cycles in nature.1,2 NH3, the primary product of nitrogen fixation, is not only a vital raw material in manufacturing, widely used in fertilizers, the military, pharmaceuticals, and other fields, but also regarded as an ideal hydrogen storage carrier due to its high energy density (3 kW h kg−1), high hydrogen capacity (17.6 wt%), ease of liquefaction, and convenience in transportation.3

Artificial nitrogen fixation dates back to the mid-19th century, but the Haber–Bosch process in 1909 ushered in its modern era.1 Since NH3, the main product of industrial nitrogen fixation, is also the primary raw material for fertilizers, the Haber–Bosch process has led to a 400% increase in world food production since its introduction, fundamentally transforming food production methods.3 However, the Haber–Bosch process requires harsh conditions of high temperature (700 K) and high pressure (100 atm) to catalyze the reaction between N2 and hydrogen (H2) to produce NH3, consuming approximately 2% of the world's total energy and emitting about 3% of global CO2 annually.3 In the Haber–Bosch process, about 72% of H2 is supplied by natural gas-based hydrogen production, and the energy consumption of the H2 production process accounts for about 75% of the total energy consumption of ammonia synthesis. The high energy consumption of the Haber–Bosch process necessitates the pursuit of greener and more sustainable pathways for artificial nitrogen fixation.

In addition to the thermocatalytic approach represented by the Haber–Bosch process, there are also enzymatic catalysis, electrocatalysis, and photocatalysis that can efficiently fix nitrogen. Microorganisms (such as bacteria) can utilize nitrogenase to reduce N2 to NH3. Another direction in microbial nitrogen fixation is the modification of nitrogen-fixing genes into eukaryotic organisms,4 such as plants, and the use of nitrogen-fixing microorganisms in electrocatalytic devices to continuously reduce N2 to NH3. These microbial nitrogen fixation processes help alleviate the demand for NH3 but are difficult to apply to modern intensive agricultural production.3 Electrocatalytic technology provides an alternative method to transfer electrons to nitrogenase without ATP hydrolysis, and various electrocatalysts have been developed to directly reduce N2 to NH3 under external electrical energy. Since the primary power generation method of China is still thermal power, utilizing renewable energy instead of electrical energy could significantly reduce energy consumption and CO2 emissions, simultaneously contributing to the goal of achieving carbon neutrality by 2060.

Photocatalytic nitrogen fixation technology, directly driven by solar energy, can catalyze the conversion of N2 to NH3 using H2O instead of H2 under mild conditions with no carbon emissions, making it a promising forward-looking technology for artificial nitrogen fixation. In 1977, Schrauzer and Guth used titanium dioxide (TiO2) as a nitrogen fixation catalyst to achieve the reduction of N2 to produce trace amounts of ammonia under illumination.5 Since then, photocatalytic nitrogen fixation technology has been receiving widespread attention. Photocatalytic nitrogen fixation technology utilizes photogenerated carriers to achieve the oxidation and reduction of reactants. The main process is illustrated in Fig. 1a.


image file: d5ra00953g-f1.tif
Fig. 1 (a) Schematic diagram of the mechanism of photocatalytic nitrogen fixation; (b) schematic diagram of N atomic orbitals and N2 molecular hybrid orbitals.6

1.1 Generation of photogenerated carriers

Photocatalysts are generally semiconductors with a valence band filled with electrons and a conduction band devoid of electrons, separated by a forbidden band. When incident light with an energy greater than or equal to the band gap (Eg) illuminates the photocatalyst, electrons in the valence band are excited to the conduction band, leaving holes in the valence band. The Eg value determines the light absorption range of the catalyst, and its maximum absorption wavelength (λ) can be calculated using the empirical formula: λ = 1240/Eg.

1.2 Separation and transport of photogenerated carriers

The photogenerated electrons and holes produced by light excitation in semiconductor photocatalysts separate from each other and migrate to the surface, but they may rapidly recombine and annihilate within picoseconds, leading to a significant portion of photogenerated carriers being unable to be effectively utilized, which significantly reduces the incident light utilization efficiency of the catalyst.7 Various strategies have been employed to promote carrier separation, including heterostructure construction,8 material compositing,9 and element doping.10

1.3 Utilization of photogenerated carriers

The electrons transported to the catalyst surface undergo reduction reactions (N2 accepts electrons to synthesize NH3), while the holes undergo oxidation reactions (H2O provides electrons to synthesize O2). N2 molecules are first enriched on the catalyst surface through physical adsorption and then activated by the catalyst, followed by reduction to the product NH3 through a continuous hydrogenation process.

To achieve efficient photocatalytic nitrogen fixation, an accurate understanding of the properties of N2 is essential. N2 is a linear molecule composed of two N atoms linked by a covalent bond (Fig. 1b). According to molecular orbital theory, the 2p orbitals of the two N atoms hybridize to form new bonding orbitals (2σ and π) and antibonding orbitals (2σ* and π*). The six electrons from the 2p orbitals fill the three bonding orbitals, forming a high-strength N[triple bond, length as m-dash]N triple bond. Thermodynamically, the N[triple bond, length as m-dash]N triple bond has a bond energy as high as 941 kJ mol−1, with the dissociation energy of the first bond reaching 410 kJ mol−1, severely hindering the dissociation of N[triple bond, length as m-dash]N; kinetically, the large energy gap (10.82 eV) of the N2 molecule obstructs electron transfer; in terms of reactivity, the low proton affinity (ΔH0 = 37.6 kJ mol−1), low electron affinity (−1.9 eV), and high ionization energy (15.85 eV) indicate that the protonation and activation of N2 are extremely difficult.11 Therefore, N2 is often used as an inert gas, and the N[triple bond, length as m-dash]N bond is considered one of the most stable chemical bonds. Due to these inherent inert properties of N2, its utilization poses a significant challenge.

Photocatalytic nitrogen fixation catalysts, as the core of photocatalytic nitrogen fixation technology, are the focus of research in this field. In recent years, several reviews have been published on nitrogen fixation photocatalyst materials12 and catalytic mechanisms,13,14 but there are fewer reviews on catalyst structures, especially structural modulation related directly to catalytic performance, such as active site structure, crystal structure, and composite structure. Therefore, this review aims to summarize the research progress in structural modulation of nitrogen fixation photocatalysts in recent years, focusing on three dimensions: active sites, crystal structure, and composite structure of nitrogen fixation photocatalysts (Fig. 2), and to propose prospects for future development, hoping to further promote the development of the field of photocatalytic nitrogen fixation.


image file: d5ra00953g-f2.tif
Fig. 2 Structural regulation of nitrogen-fixing photocatalysts.

2. Structural construction of photocatalytic nitrogen fixation catalysts

2.1 Construction of active sites

The design and construction of active sites are crucial for achieving efficient photocatalytic nitrogen fixation. Active sites need not only to effectively adsorb N2 but also to rapidly accept electrons and protons for N2 reduction. Currently, the performance of nitrogen fixation photocatalysts still lags significantly behind practical production demands. On the one hand, the difficulty in activating N2 limits the activity of photocatalytic nitrogen fixation. On the other hand, competition from hydrogen evolution reactions constrains the selectivity of photocatalytic nitrogen fixation. The activation mechanism of N2 is generally accepted as the π-backdonation mechanism proposed by Han et al.,13 which divides N2 activation into two processes: process 1 involves the empty d-orbitals of the metal accepting electrons from the σg bonding orbital of N2, and process 2 involves the occupied d-orbitals of the metal donating electrons to the image file: d5ra00953g-t1.tif antibonding orbital of N2. It can be seen that N2 activation is essentially an electron exchange process between the active site and the N2 molecule. Therefore, enhancing the electron exchange between the active site and the N2 molecule is an effective way to promote N2 activation. On the one hand, the transfer of electrons from N2 to the metal can be enhanced by pulling the two lone pairs of electrons at the ends of N2. Wang et al.15 anchored B atoms on the surface of metal-free black phosphorus materials based on first-principles theoretical calculations, where adjacent B atoms formed “Lewis acid pairs” generating a “pull–pull effect” on N2 molecules. Simulation calculations showed that the B atoms could attract electrons from the N2 molecules and transfer them to the B atoms, and the resulting “pull–pull effect” could reduce the activation energy barrier of N2 and promote N2 activation to a certain extent. Building on this research, Wang et al.16 further used bimetallic sites instead of monometallic sites to promote the transfer of electrons from N2 to the metal. They screened FeMo/g-C3N4, TiMo/g-C3N4, MoW/g-C3N4, and NiMo/g-C3N4 catalysts among 28 catalyst combinations through theoretical calculations. These four catalysts have suitable bandgap structures and visible light absorption capabilities, making them promising candidates for efficient nitrogen fixation photocatalysts. On the other hand, methods such as constructing heterojunctions17 and forming localized surface plasmon resonance fields15 can be used to promote the transfer of electrons from metal to N2. Huang et al.17 loaded Bi metal cocatalysts on the surface of BiOBr to form a Bi/semiconductor interfacial Schottky junction. The introduction of Bi could not only promote interfacial electron transfer in the photocatalyst but also facilitate N2 activation due to its strong electron donation. This unidirectional electron transfer from BiOBr to the active site Bi increased the efficiency of photocatalytic ammonia synthesis by 65 times. Bimetallic systems exhibit enhanced nitrogen activation by decoupling the electron acceptance and donation processes. For instance, Zhao et al.18 demonstrated that bimetallic organic frameworks (BMOFs) composed of hard-acid (high ionization potential, In) and soft-acid (low In) metal nodes synergistically promote N[triple bond, length as m-dash]N bond cleavage. The hard-acid metal (e.g., Fe) facilitates electron withdrawal from N2 via σ-bond interaction, while the soft-acid metal (e.g., Sr) donates electrons to the π*-antibonding orbitals of N2. This decoupling strategy reduces the activation energy barrier and enhances electron exchange efficiency, achieving an NH3-evolution rate of 780 μmol g−1 h−1 under visible light.

In recent years, constructing coordination unsaturated sites through vacancies has become one of the research hotspots in the field of photocatalytic nitrogen fixation. Vacancies in photocatalysts can not only promote the adsorption of N2 but also facilitate charge separation to a certain extent. Vacancies are mainly classified into oxygen vacancies, nitrogen vacancies, carbon vacancies, and sulfur vacancies based on the missing elements, with oxygen vacancies being widely used in the field of photocatalytic nitrogen fixation. Shiraishi et al.19 investigated the effect of oxygen vacancies on the photocatalytic nitrogen fixation of TiO2. They proposed that Ti3+ groups formed at the oxygen vacancies on the TiO2 surface served as active sites that could effectively capture electrons and promote N2 activation (Fig. 3a). To better create oxygen vacancies, researchers typically choose materials rich in oxygen elements as photocatalysts. Layered double hydroxides (LDH) are increasingly gaining attention due to their controllable electronic structure and low cost. Zhang et al.20 were the first to apply CuCr-LDH nanosheets (CuCr-NS) rich in oxygen vacancies to the field of photocatalytic nitrogen fixation. Under full-spectrum irradiation, the photocatalytic ammonia synthesis rate of CuCr-NS was 78.6 μmol g−1 h−1, with an apparent quantum efficiency (AQE) of 2.4% at a wavelength of 400 nm. Subsequently, they optimized the oxygen vacancies by incorporating coordination unsaturated Cuδ+ species onto ZnAl-LDH nanosheets (Fig. 3b).21 DFT theoretical calculations indicated that the oxygen vacancies and Cuδ+ species could effectively promote N2 adsorption and activation.


image file: d5ra00953g-f3.tif
Fig. 3 (a) Schematic diagram of the photocatalytic nitrogen fixation cycle on the oxygen vacancies of TiO2;20 (b) schematic diagram of the photocatalytic nitrogen fixation process of ZnAl-LDH nanosheets and the formation process of electron-rich Cuδ+ and oxygen vacancies.21

Zhang et al.22 reported BiOBr nanosheets (BOB-001-OV) with oxygen vacancies that could stretch the N[triple bond, length as m-dash]N bond length from 1.078 Å to 1.133 Å. Theoretical calculations showed that the defect states induced by oxygen vacancies could act as electron acceptors to effectively inhibit electron–hole pair recombination and enhance charge transfer from BOB-001-OV to N2 molecules. Fluorescence spectroscopy revealed that the average lifetime (τ) of BOB-001-OV was 2.15 ns, approximately twice that of BiOBr without oxygen vacancies, demonstrating that oxygen vacancies could promote the migration of photogenerated carriers. The ammonia synthesis rate of BOB-001-OV under visible light was 104.3 μmol g−1 h−1, with an AQE of 0.23% at 420 nm. Simultaneously, the generated O2 was stoichiometrically close to 75% of produced NH3, proving that H2O could serve as an electron donor to achieve a complete chemical cycle in photocatalytic nitrogen fixation.

Single-atom catalysts (SACs) have garnered significant attention due to their rapid electron–hole separation and targeted active sites. The atomic-level dispersion of SACs maximizes atomic utilization and can anchor SACs as active sites to enhance nitrogen fixation activity in the field of photocatalytic nitrogen fixation. Wang et al.23 established a g-C3N4 model anchored with single B atoms (B/g-C3N4) and simulated the reduction process of N2 on B/g-C3N4 through first-principles calculations (Fig. 4a). The calculation results showed that B/g-C3N4 could effectively reduce gaseous N2 molecules through an enzymatic mechanism at an extremely low initial potential. In addition, anchoring single B atoms significantly enhanced the visible and infrared light absorption of g-C3N4, granting the catalyst strong N2 reduction capabilities. To address the issue of insufficient local electron supply. Yin et al.25 prepared the Ru-SA/HxMoO3−y photocatalyst by combining Ru single atoms (Ru-SA) with molybdenum oxide rich in oxygen vacancies and conducted photocatalytic nitrogen fixation studies using N2 and H2 as reactants (Fig. 4b). Ru-SA facilitated the activation of N2 molecules and the migration of H2 molecules, while Mo sites containing oxygen vacancies captured local electrons and catalyzed the reduction of N2 molecules, ultimately achieving an ammonia synthesis rate of 4 mmol g−1 h−1 for SA/HxMoO3−y.


image file: d5ra00953g-f4.tif
Fig. 4 (a) Design concept of the B/g-C3N4 photocatalyst;23 (b) schematic diagram of the structure and preparation process of the Ru-SA/HxMoO3−y photocatalyst.24

2.2 Crystal structure construction

Research on the regulation of the crystal structure of photocatalysts for photocatalytic nitrogen fixation primarily focuses on crystal form, crystal facet and morphology control. Semiconductor photocatalysts are mostly crystalline materials, and their orderly crystal structure facilitates electron transfer and reduces the recombination of photogenerated carriers. Crystal form is a fundamental characteristic of crystalline materials, and research into the crystal form of nitrogen fixation photocatalysts can be traced back to the inception of photocatalytic nitrogen fixation. In 1977, Schrauzer et al.5 found that after calcination at 1000 °C, some TiO2 impregnated with iron sulfate could transform from the anatase phase to the rutile phase, achieving the highest photocatalytic nitrogen fixation activity when the loading of Fe2O3 reached 0.2%. Active site density is one of the critical factors directly affecting photocatalytic efficiency. Overly ordered crystal structures cannot expose sufficient active sites for N2 adsorption, thereby hindering the adsorption and activation of N2 molecules on the active sites to some extent. Hou et al.24 prepared amorphous SmOCl nanosheet materials using graphene oxide as a template. With increasing amorphization of the SmOCl nanosheets, numerous oxygen vacancies formed on the surface of the nanosheets, significantly promoting the adsorption and activation of N2 molecules. Additionally, the authors observed enhanced Sm–O covalent bonds, indicating electron migration from bulk to surface for combination with N2. Under illumination, the ammonia synthesis rate of SmOCl nanosheets reached 426 μmol g−1 h−1, with an AQE of 0.32% at 420 nm.

By precisely and rationally exposing specific crystal facets, the crystals with target facets can be controlled at the atomic scale. The atomic distribution on the crystal facets also significantly influences the active sites and electronic structure of photocatalysts, prompting an increasing number of researchers to attempt to enhance photocatalytic nitrogen fixation activity by regulating the crystal facets of semiconductors. Atoms arrange differently on different exposed facets, hence facet regulation can significantly impact the enhancement of photocatalytic nitrogen fixation activity. Bai et al.27 prepared two types of Bi5O7I nanosheets exposing the (100) and (001) facets through hydrolysis and calcination methods. Further experiments demonstrated that the conduction band potential of Bi5O7I-001 (−1.45 V) was more negative than that of Bi5O7I-100 (−0.85 V), and Bi5O7I-001 exhibited higher photogenerated carrier separation efficiency and photocatalytic ammonia synthesis activity. When using ethanol as a hole scavenger, the photocatalytic ammonia synthesis rate of Bi5O7I-001 reached 111.5 μmol g−1 h−1. Li et al.26 systematically investigated the influence of the (001) and (010) facets of BiOCl nanosheets on the adsorption and activation of N2 (Fig. 5a). Experimental results showed that N2 molecules primarily bind through a terminal pathway on (001) facet of BiOCl (Fig. 5b), whereas they primarily bind through a side-on bridging pathway on (010) facet (Fig. 5c). The photocatalytic ammonia synthesis rate of (010) facet of BiOCl was approximately 2.5 times higher than that of (001) facet, reaching 4.62 μmol g−1 h−1. Zhang et al.28 combined facet regulation with defect regulation by constructing defects on the (040) facet of BiVO4 single crystals, which induced V4+/V5+ sites. Among them, V4+ sites are responsible for the chemical adsorption of N2, while V5+ sites serve as bridges for electron transfer, delivering electrons captured by defects to V4+ sites for NH3 synthesis. In contrast, the (110) facet of BiVO4 only contains V5+ sites. Therefore, the authors regulated the ratio of (040)/(110) facets by changing the pH during catalyst preparation, achieving a photocatalytic ammonia synthesis rate of 103.4 μmol g−1 h−1.


image file: d5ra00953g-f5.tif
Fig. 5 (a) Crystal structure of BiOCl and corresponding (001) and (010) crystal facets; (b) N2 molecules bind to the (001) crystal facet of BiOCl through a terminal pathway; (c) N2 molecules bind to the (010) crystal facet of BiOCl through a side-on bridging pathway.26

Controllably regulating the morphology of materials is also an effective strategy for crystal regulation, such as one-dimensional nanowires/nanofibers/nanorods, two-dimensional ultrathin nanosheets, and three-dimensional porous/hollow structures. One-dimensional nanostructures can provide direct electron transfer channels, facilitate photogenerated electron transfer, and possess a large specific surface area, providing sufficient active sites for N2 adsorption. Sun et al.29 applied Nb2O5 nanofibers to the field of nitrogen fixation, benefiting from rapid electron exchange and transfer between Nb atoms and N2 molecules, which led to rapid polarization and activation of adsorbed N2 molecules. Compared to one-dimensional materials, ultrathin two-dimensional nanosheets feature large lateral dimensions and specific surface areas, which are highly advantageous for the adsorption of N2 molecules on the surface of photocatalyst. Meanwhile, the high exposure of atoms on two-dimensional nanosheet surfaces suits them for surface modification and functionalization. This includes element doping and vacancy creation, enhancing N2 adsorption and activation on the catalyst. Cao et al.30 constructed a MXene-Based 2D/2D Ti3C2/TiO2 heterojunction with spatially separated redox sites for photocatalytic N2 reduction reaction. Electron-rich unsaturated Ti sites on Ti3C2 MXene are instrumental in adsorption and activation of N2, exhibiting a high NH3 production rate of 24.4 μmol g−1 h−1. Three-dimensional materials integrate and optimize the structural advantages of two-dimensional materials. For example, porosity allows rapid transport of N2 molecules and protons, while hollow structures can utilize internal cavities to reflect and refract incident light for multiple utilizations. Furthermore, three-dimensional flower-like nanospheres assembled from two-dimensional nanosheets combine the advantages of both two-dimensional and three-dimensional materials, potentially boosting light absorption and N2 adsorption synergistically. Zhang et al.31 reported that Ru-decorated MoO3−x microspheres exhibit dynamic oxygen vacancies which cycle between N2 adsorption (Ru/MoO3−x) and NH3 desorption (Ru/MoO3−xNy) states. This H*-mediated oxygen vacancies evolution strategy achieved an NH3 production rate of 192.38 μmol g−1 h−1, 2.68-fold higher than that of pristine MoO3−x. The three-dimensional framework promotes light scattering and provides interconnected pathways for charge transport, addressing the limitations of 2D materials in carrier recombination. However, two-dimensional materials and three-dimensional materials still exist several limitations. Two-dimensional materials (such as ultra-thin nanosheets) present structural instability due to high surface energy, and excessive lateral size may hinder mass transfer. Although three-dimensional porous materials can optimize light absorption, their complex synthesis process limits their large-scale application.32

2.3 Construction of composite structures

Composite catalysts are comprised of multiple components. Compared to single structures, their composite structures can better overcome the inherent shortcomings of each component. Even more, the interplay between components may create synergistic effects, achieving a “1 + 1 > 2” outcome. Heterostructures are typical composite structures that regulate photoelectric properties by combining two semiconductor materials with different band-gap structures. Ghosh et al.33 embedded Bi2MoO6 into g-C3N4 nanosheets to obtain a g-C3N4/Bi2MoO6 heterojunction, which exhibited efficient photogenerated carrier separation and N2 reduction capabilities. Liu et al.10 used TiO2@C combined with g-C3N4 to obtain the TiO2@C/g-C3N4 composite catalyst. Compared to bulk g-C3N4, the steady-state photoluminescence spectrum peak intensity of TiO2@C/g-C3N4 significantly decreased. Additionally, the transient photocurrent intensity of TiO2@C/g-C3N4 was approximately 1.7 times higher than that of bulk g-C3N4, further proving the positive effect of composite structures on promoting carrier separation. When the molar ratio of TiO2@C to g-C3N4 was 10[thin space (1/6-em)]:[thin space (1/6-em)]1, the photocatalytic ammonia synthesis rate of TiO2@C/g-C3N4 reached 250.6 μmol g−1 h−1. Xia et al.34 utilized the intrinsic electric field differences between Bi2O3 and CoAl-LDH to construct Bi2O3@CoAl-LDHs core–shell heterojunction photocatalysts for nitrogen fixation. Hollow Bi2O3 microspheres were first prepared using a solvothermal method and then mixed with the precursor solution of CoAl-LDH for a secondary solvothermal reaction to obtain the core–shell composite Bi2O3@CoAl-LDHs (Fig. 6). This composite structure photocatalyst had a large interface contact area, effectively promoting carrier separation and achieving a photocatalytic ammonia synthesis rate of 48.7 μmol g−1 h−1.
image file: d5ra00953g-f6.tif
Fig. 6 Schematic diagram of the preparation route of Bi2O3@CoAl-LDHs.34

Apart from compositing with semiconductor materials, metal compositing is also a common strategy to promote photogenerated electron–hole separation. In particular, noble metal loading can form Schottky barriers with semiconductors. The resulting energy band bending can effectively inhibit the recombination of photogenerated carriers. Ranjit et al.35 loaded Ru, Rh, Pd, and Pt onto TiO2 for photocatalytic nitrogen fixation. The authors found that the photocatalytic activity of metal-loaded TiO2, in descending order, was: Ru > Rh > Pd > Pt. The highest performance was observed for Ru-loaded TiO2 due to its higher Ru–H bond strength. Ye et al.36 prepared the Ni2P/Cd0.5Zn0.5S composite photocatalyst by loading transition metal phosphide Ni2P onto Cd0.5Zn0.5S. The authors demonstrated through time-resolved PL spectroscopy, photocurrent, and electrochemical impedance spectroscopy that Ni2P/Cd0.5Zn0.5S had higher photogenerated electron–hole separation efficiency compared to Ni2P and Cd0.5Zn0.5S. Maimaitizi et al.37 synthesized flower-like N-MoS2 microspheres through a one-step solvothermal method and then prepared Pt/N-MoS2 photocatalysts via a photo-ultrasonic reduction method. The special multi-level flower-like structure of Pt/N-MoS2 could provide more active sites. N doping could narrow the bandgap of MoS2, enhancing its response to visible light. Pt nanoparticles could act as electron traps, forming Schottky barriers, thereby improving photogenerated carrier separation efficiency. They demonstrated that the nitrogen fixation activity of ultrasonic photocatalysis was higher than that of ultrasonic catalysis and photocatalysis, indicating a synergistic effect between ultrasound and visible light irradiation. Schrauzer et al.5 studied the impact of Fe, Cr, Co, and Mo metals on the photocatalytic nitrogen fixation performance of TiO2. They found that Fe metal significantly enhanced the photocatalytic nitrogen fixation activity. Based on this, Zhao et al.38 further investigated the effect of Fe loading on TiO2 photocatalytic nitrogen fixation and found that an appropriate concentration of Fe3+ could act as a hole trapping site to inhibit the recombination of photogenerated carriers, thereby enhancing photocatalytic activity. Liu et al.39 also reported on Fe-loaded SrMoO4 photocatalysts for nitrogen fixation reactions. They found that as the Fe loading increased from 0 to 5.1%, the bandgap of SrMoO4 decreased from 3.98 eV to 2.93 eV. The bandgap directly affects the light absorption range, so the reduction in band-gap expanded the light absorption range of SrMoO4 from the UV region to the visible region. Additionally, Fe loading could induce surface defects, which could serve as N2 adsorption sites and inhibit electron–hole pair recombination, thereby effectively enhancing the photocatalytic nitrogen fixation efficiency. Compared to single-metal loading, bimetals may exhibit more effective synergistic effects during photocatalytic nitrogen fixation. Li et al.40 loaded Pt-doped Fe nanoclusters onto g-C3N4. This study showed that Pt–Fe nanoclusters could cause upward bending of the semiconductor energy band and form a large Schottky barrier, thereby enhancing the separation of photogenerated carriers and N2 reduction. Therefore, the FePt@C3N4 photocatalyst exhibited an ammonia synthesis rate of 3.7 μmol g−1 h−1 under visible light irradiation. In addition to photocatalytic nitrogen fixation performance, structural regulation may also impact other properties of the catalyst. Zhang et al.41 demonstrated that introducing oxygen vacancies into Ru/W18O49 significantly enhanced the hydrogen spillover effect, which not only improved ammonia synthesis activity but also optimized the hydrogen adsorption/desorption thermodynamics and charge separation efficiency. Specifically, the Schottky junction formed between Ru and W18O49 promoted electron trapping at Ru sites, reducing recombination losses. Additionally, the OVs-rich structure improved long-term stability, retaining 87% activity after six cycles, highlighting the dual role of structural defects in enhancing both efficiency and durability.

3. Summary and outlook

Photocatalytic nitrogen fixation, as a green and sustainable new process for ammonia synthesis, holds broad application prospects. This paper first briefly introduces the basic principles and existing issues of photocatalytic nitrogen fixation, followed by a review focusing on three dimensions of active site, crystal structure, and composite structure construction in nitrogen fixation photocatalysts. In terms of active site construction, active site design can be approached from the mechanism of photocatalytic nitrogen fixation, strengthening the electron exchange process in nitrogen activation, thereby enhancing the efficiency of photocatalytic nitrogen fixation. Additionally, holes and single atoms are also important research directions. In terms of crystal structure construction, current research primarily focuses on crystal polymorphism, crystal facet, and morphology regulation. High-performance crystal facets can be specifically exposed to enhance activity. Meanwhile, special morphologies such as flower-sphere, hollow microspheres, and core–shell structures are conducive to the reflection and refraction of incident light, increasing the utilization efficiency of incident light. In composite structure construction, both Schottky barriers formed by semiconductor–metal combinations and heterojunctions formed by semiconductor–semiconductor combinations promote photogenerated carrier separation, yielding synergistic effects for “1 + 1 > 2” outcomes. Finally, based on the current research progress in the field of photocatalytic nitrogen fixation, this paper provides an outlook for the field, hoping to promote its development:

(1) Rational design of ternary transfer channels for molecules, protons, and electrons. Photocatalytic nitrogen fixation is a typical complex reaction requiring the simultaneous participation of N2 molecules, protons, and electrons. The transfer paths, transfer rates, and transfer mechanisms for these three differ. N2 molecules primarily diffuse with the aqueous solution, protons primarily transfer along the catalyst surface, and electrons primarily transfer along the catalyst framework. Direct contact between protons and electrons may lead to hydrogen evolution reactions. The design of respective transfer channels for molecules, protons, and electrons must ensure rapid material transfer while maintaining independence between channels, posing significant challenges for the structural design of photocatalysts.

(2) Exploration of new mechanisms for photocatalytic nitrogen fixation. Although the photocatalysts prepared in existing research have achieved high activity and selectivity, there is still a significant gap compared to biological nitrogenase. The rich species diversity in nature encompasses various efficient synergistic mechanisms, such as confinement effects, channel effects, compartmentalization effects, and proximity effects. If we can learn from natural coordination mechanisms and reveal new mechanisms for photocatalytic nitrogen fixation using interdisciplinary approaches such as materials genomics and artificial intelligence, it will bring new developments to this field.

(3) Expansion of photocatalyst applications. Photocatalytic nitrogen fixation reactions utilize photogenerated electrons to reduce N2 molecules. To facilitate the analysis of the catalytic process, hole scavengers are often used to consume photogenerated holes. However, photogenerated holes possess strong oxidizing capabilities. If holes can be utilized to catalyze oxidation to produce higher-value chemicals, it will further broaden the application prospects of photocatalytic technology. Additionally, the stability test time for photocatalysts is generally within 24 hours, and the photocatalytic ammonia synthesis rate is generally at the micromolar level, still far from practical application requirements. Developing catalytic materials and optimizing structural parameters will be key factors in expanding the application of photocatalytic nitrogen fixation.

Data availability

No primary research results, software or code have been included and no new data were generated or analysed as part of this review.

Conflicts of interest

There are no conflicts of interest to declare.

References

  1. A. J. Medford and M. C. Hatzell, Photon-driven nitrogen fixation: Current progress, thermodynamic considerations, and future outlook, ACS Catal., 2017, 7, 2624–2643 CrossRef CAS.
  2. A. Chen and B. Y. Xia, Ambient dinitrogen electrocatalytic reduction for ammonia synthesis, J. Mater. Chem. A, 2019, 7, 23416–23431 RSC.
  3. J. G. Chen, R. M. Crooks, L. C. Seefeldt, K. L. Bren, R. M. Bullock, M. Y. Darensbourg, P. L. Holland, B. Hoffman, M. J. Janik, A. K. Jones, M. G. Kanatzidis, P. King, K. M. Lancaster, S. V. Lymar, P. Pfromm, W. F. Schneider and R. R. Schrock, Beyond fossil fuel-driven nitrogen transformations, Science, 2018, 360, eaar6611 CrossRef PubMed.
  4. E. J. Vicente and D. R. Dean, Keeping the nitrogen-fixation dream alive, Proc. Natl. Acad. Sci. U. S. A., 2017, 114, 3009–3011 CrossRef CAS PubMed.
  5. G. N. Schrauzer and T. D. Guth, Photolysis of water and photoreduction of nitrogen on titanium dioxide, J. Am. Chem. Soc., 1977, 99, 7189–7193 CrossRef CAS.
  6. S. Y. Wang, F. Ichihara, H. Pang, H. Chen and J. H. Ye, Nitrogen fixation reaction derived from nanostructured catalytic materials, Adv. Funct. Mater., 2018, 28, 1803309 CrossRef.
  7. J. Schneider, M. Matsuoka, M. Takeuchi, J. Zhang, Y. Horiuchi, M. Anpo and D. W. Bahnemann, Understanding TiO2 photocatalysis: Mechanisms and materials, Chem. Rev., 2014, 114, 9919–9986 CrossRef CAS.
  8. Y. Han, X. Lu, S. Tang, X. Yin, Z. Wei and T. Lu, Metal-free 2D/2D heterojunction of graphitic carbon nitride/graphdiyne for improving the hole mobility of graphitic carbon nitride, Adv. Energy Mater., 2018, 8, 1702992 CrossRef.
  9. H. Liang, J. Li and Y. Tian, Construction of full-spectrum-driven Ag-g-C3N4/W18O49 heterojunction catalyst with outstanding N2 photofixation ability, RSC Adv., 2017, 7, 42997–43004 RSC.
  10. Q. X. Liu, L. H. Ai and J. Jiang, MXene-derived TiO2@C/g-C3N4 heterojunctions for highly efficient nitrogen photofixation, J. Mater. Chem. A, 2018, 6, 4102–4110 Search PubMed.
  11. X. Chen, N. Li, Z. Kong, W.-J. Ong and X. Zhao, Photocatalytic fixation of nitrogen to ammonia: State-of-the-art advancements and future prospects, Mater. Horiz., 2018, 5, 9–27 Search PubMed.
  12. S. Lin, X. Zhang, L. Chen, Q. Zhang, L. Ma and J. Liu, A review on catalysts for electrocatalytic and photocatalytic reduction of N2 to ammonia, Green Chem., 2022, 24, 9003–9026 Search PubMed.
  13. Q. Han, H. Jiao, L. Xiong and J. Tang, Progress and challenges in photocatalytic ammonia synthesis, Mater. Adv., 2021, 2, 564–581 Search PubMed.
  14. P. Praus, Photocatalytic Nitrogen Fixation using Graphitic Carbon Nitride: A Review, ChemistrySelect, 2023, 8, e202204511 Search PubMed.
  15. L. Shi, Q. Li, C. Ling, Y. Zhang, Y. Ouyang, X. Bai and J. Wang, Metal-free electrocatalyst for reducing nitrogen to ammonia using a Lewis acid pair, J. Mater. Chem. A, 2019, 7, 4865–4871 RSC.
  16. S. Y. Wang, L. Shi, X. W. Bai, Q. Li, C. Y. Ling and J. L. Wang, Highly efficient photo-/electrocatalytic reduction of nitrogen into ammonia by dual-metal sites, ACS Cent. Sci., 2020, 6, 1762–1771 Search PubMed.
  17. Y. Huang, Y. Zhu, S. Chen, X. Xie, Z. Wu and N. Zhang, Schottky junctions with Bi cocatalyst for taming aqueous phase N2 reduction toward enhanced solar ammonia production, Adv. Sci., 2021, 8, 2003626 Search PubMed.
  18. Z. Zhao, H. Ren, D. Yang, Y. Han, J. Shi, K. An, Y. Chen, Y. Shi, W. Wang, J. Tan, X. Xin, Y. Zhang and Z. Jiang, Boosting nitrogen activation via bimetallic organic frameworks for photocatalytic ammonia synthesis, ACS Catal., 2021, 11, 9986–9995 CrossRef CAS.
  19. H. Hirakawa, M. Hashimoto, Y. Shiraishi and T. Hirai, Photocatalytic conversion of nitrogen to ammonia with water on surface oxygen vacancies of titanium dioxide, J. Am. Chem. Soc., 2017, 139, 10929–10936 CrossRef CAS PubMed.
  20. Y. Zhao, X. Jia, G. I. N. Waterhouse, L. Wu, C. Tung, D. O'hare and T. Zhang, Layered double hydroxide nanostructured photocatalysts for renewable energy production, Adv. Energy Mater., 2016, 6, 1501974 Search PubMed.
  21. S. Zhang, Y. Zhao, R. Shi, C. Zhou, G. I. N. Waterhouse, L. Wu, C.-H. Tung and T. Zhang, Efficient photocatalytic nitrogen fixation over Cuδ+-modified defective ZnAl-layered double hydroxide nanosheets, Adv. Energy Mater., 2020, 10, 1901973 Search PubMed.
  22. H. Li, J. Shang, Z. Ai and L. Zhang, Efficient visible light nitrogen fixation with BiOBr nanosheets of oxygen vacancies on the exposed {001} facets, J. Am. Chem. Soc., 2015, 137, 6393–6399 CrossRef CAS PubMed.
  23. C. Ling, X. Niu, Q. Li, A. Du and J. Wang, Metal-free single atom catalyst for N2 fixation driven by visible light, J. Am. Chem. Soc., 2018, 140, 14161–14168 CrossRef CAS PubMed.
  24. H. Yin, Z. Chen, Y. Peng, S. Xiong, Y. Li, H. Yamashita and J. Li, 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(14), e202114242 CrossRef CAS PubMed.
  25. H. Yin, Z. Chen, Y. Peng, S. Xiong, Y. Li, H. Yamashita and J. Li, 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 CrossRef CAS PubMed.
  26. H. Li, J. Shang, J. Shi, K. Zhao and L. Zhang, Facet-dependent solar ammonia synthesis of BiOCl nanosheets via a proton-assisted electron transfer pathway, Nanoscale, 2016, 8, 1986–1993 RSC.
  27. Y. Bai, L. Ye, T. Chen, L. Wang, X. Shi, X. Zhang and D. Chen, Facet-dependent photocatalytic N2 fixation of bismuth-rich Bi5O7I nanosheets, ACS Appl. Mater. Interfaces, 2016, 8, 27661–27668 CrossRef CAS PubMed.
  28. G. H. Zhang, Y. Meng, B. Xie, Z. M. Ni, H. F. Lu and S. J. Xia, Precise location and regulation of active sites for highly efficient photocatalytic synthesis of ammonia by facet-dependent BiVO4 single crystals, Appl. Catal., B, 2021, 296, 120379 CrossRef CAS.
  29. J. Han, Z. Liu, Y. Ma, G. Cui, F. Xie, F. Wang, Y. Wu, S. Gao, Y. Xu and X. Sun, Ambient N2 fixation to NH3 at ambient conditions: Using Nb2O5 nanofiber as a high-performance electrocatalyst, Nano Energy, 2018, 52, 264–270 CrossRef CAS.
  30. C. Cao, J. Li, L. Zhang, Y. Hu, L. Zhang and W. Yang, MXene-Based 2D/2D Ti3C2/TiO2 Heterojunction with Spatially Separated Redox Sites for Efficient Photocatalytic N2 Reduction towards NH3, J. Mater. Sci. Technol., 2025, 214, 180–193 CrossRef CAS.
  31. L. Zhang, R. Li, L. Guo, L. Cui, X. Zhang, Y. Wang, Y. Wang, X. Jian, X. Gao, C. Fan, J. Wang and J. Liu, Active Hydrogen-Switchable Dynamic Oxygen Vacancies in MoO3–x upon Ru Nanoparticle Decoration for Boosting Photocatalytic Ammonia Synthesis Performance, ACS Catal., 2024, 14, 5696–5709 CrossRef CAS.
  32. Y. Xia, Y. Xu, X. Yu, K. Chang, H. Gong, X. Fan, X. Meng, X. Huang, T. Wang and J. He, Structural Design and Control of Photocatalytic Nitrogen-Fixing Catalysts, J. Mater. Chem. A, 2022, 10, 17377–17394 RSC.
  33. V. E. Kermani, H. A. Yangjeh, D. H. Khalilabad and S. Ghosh, Nitrogen photofixation ability of g-C3N4 nanosheets/Bi2MoO6 heterojunction photocatalyst under visible-light illumination, J. Colloid Interface Sci., 2020, 563, 81–91 CrossRef PubMed.
  34. S. Xia, G. Zhang, Z. Gao, Y. Meng, B. Xie, H. Lu and Z. Ni, 3D hollow Bi2O3@CoAl-LDHs direct Z-scheme heterostructure for visible-light-driven photocatalytic ammonia synthesis, J. Colloid Interface Sci., 2021, 604, 798–809 CrossRef CAS PubMed.
  35. K. T. Ranjit, T. K. Varadarajan and B. Viswanathan, Photocatalytic reduction of dinitrogen to ammonia over noble-metal-loaded TiO2, J. Photochem. Photobiol., A, 1996, 96, 181–185 CrossRef CAS.
  36. L. Ye, C. Han, Z. Ma, Y. Leng, J. Li, X. Ji, D. Bi, H. Xie and Z. Huang, Niloading on Cd0.5Zn0.5S solid solution for exceptional hotocatalytic nitrogen fixation under visible light, Chem. Eng. J., 2017, 307, 311–318 CrossRef CAS.
  37. H. Maimaitizi, A. Abulizi, T. Zhang, K. Okitsu and J. J. Zhu, Facile photo-ultrasonic assisted synthesis of flower-like Pt/N-MoS2 microsphere as an efficient sonophotocatalyst for nitrogen fixation, Ultrason. Sonochem., 2020, 63, 104956 CrossRef CAS PubMed.
  38. W. Zhao, J. Zhang, X. Zhu, M. Zhang, J. Tang, M. Tan and Y. Wang, Enhanced nitrogen photofixation on Fe-doped TiO2 with highly exposed (101) facets in the presence of ethanol as scavenger, Appl. Catal., B, 2014, 144, 468–477 CrossRef CAS.
  39. J. Luo, X. Bai, Q. Li, X. Yu, C. Li, Z. Wang, W. Wu, Y. Liang, Z. Zhao and H. Liu, Band structure engineering of bioinspired Fe doped SrMoO4 for enhanced photocatalytic nitrogen reduction performance, Nano Energy, 2019, 66, 104187 CrossRef CAS.
  40. Z. Li, Z. Gao, B. Li, L. Zhang, R. Fu, Y. Li, X. Mu and L. Li, Fe-Pt nanoclusters modified Mott-Schottky photocatalysts for enhanced ammonia synthesis at ambient conditions, Appl. Catal., B, 2020, 262, 118276 CrossRef CAS.
  41. L. Zhang, R. Li, L. Cui, Z. Sun, L. Guo, X. Zhang, Y. Wang, Y. Wang, Z. Yu, T. Lei, X. Jian, X. Gao, C. Fan and J. Liu, Boosting Photocatalytic Ammonia Synthesis Performance Over OVs-Rich Ru/W18O49: Insights into The Roles of Oxygen Vacancies in Enhanced Hydrogen Spillover Effect, Chem. Eng. J., 2023, 461, 141892 CrossRef CAS.

Footnote

Zhao Zhanfeng and Zhang Yue contributed equally to the work.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.