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

Research progress in photochemical synthesis of urea through C–N coupling reactions

Zhidong Yang* and Peixia Li
School of Chemistry and Environment, Ankang University, Ankang 725000, Shanxi, China. E-mail: yangzd1618@163.com

Received 24th October 2025 , Accepted 3rd December 2025

First published on 9th December 2025


Abstract

Urea, as an important organic compound, plays significant role in promoting the development of agriculture, industry and biological sciences. Conventional urea synthesis process requires harsh reaction conditions (high temperature and pressure), making it energy-intensive, emission-intensive, and costly. Photocatalytic C–N coupling is a potentially green and promising alternative strategy for synthesizing valuable urea from CO2 and inexpensive nitrogen sources (such as N2, nitrates, ammonia) as feedstocks under ambient conditions using solar energy. However, the specific details of urea photosynthesis have not been systematically reviewed so far. This article reviews the basic principles of photocatalytic urea synthesis, including the fundamentals, key intermediates, product identification and quantification. Meanwhile, it comprehensively summarizes research advances in photocatalytic urea synthesis, with a focus on the application and design principles of photocatalysts in urea production. Finally, the major challenges and prospective research directions in the field of photocatalytic urea synthesis are thoroughly discussed. We hope that this review will provide useful insights to inspire future research and discoveries.


1 Introduction

Urea is a compound with critical applications across multiple fields.1,2 In agriculture, it serves as an efficient nitrogen fertilizer that improves soil quality, promotes plant growth, and enhances crop yield and quality.3,4 In industry, it is an important raw material for fine chemicals such as plastics, resins, coatings, adhesives, pigments, and dyes.5,6 In the energy sector, urea can be used as a proton conductor in fuel cells to improve performance, and it can also produce hydrogen through pyrolysis or electrolysis, serving as a potential hydrogen source.5 In medicine, urea exhibits functions such as keratin softening, wound healing, antipruritic, and antibacterial effects, making it widely used in dermatological treatments.7,8 Therefore, the development of the urea synthesis industry is of great significance for ensuring global food security, advancing chemical technology, promoting economic growth, and improving human living standards.9

The industrial urea synthesis process involves the reaction of ammonia and carbon dioxide under harsh conditions to produce ammonium carbamate, which is then dehydrated to produce urea.10–13 As one of the reaction feedstocks for urea synthesis, approximately 80% of the ammonia is derived from the Haber–Bosch process.14 The Haber–Bosch process requires high-temperature and high-pressure conditions (400–650 °C and 200–400 bar), accounting for 1–2% of global energy consumption.14–16 Hydrogen, as another key reactant in the Haber–Bosch process, is typically produced via steam methane reforming (CH4 + 2H2O → 4H2 + CO2) or coal steam reaction (C + 2H2O → 2H2 + CO2), both of which generate significant CO2 emissions.17 It is estimated that the Haber–Bosch process contributes to about 1.5% of global anthropogenic CO2 emissions.17 As a result, conventional urea synthesis not only consumes substantial energy but also exacerbates environmental burdens. It is urgent to explore more environmentally friendly and energy-saving urea production processes and technologies to reduce environmental pollution and resource consumption.18

Fortunately, solar-driven photosynthesis techniques are able to initiate chemical reactions under mild conditions, thereby reducing energy consumption and environmental issue and offering prospects for sustainable development and green chemistry.19 Photocatalytic C–N coupling reactions can convert carbon species such as CO2, CO, methanol and nitrogen species such as N2, nitrite, nitrate, or NO into valuable urea under ambient conditions and in aqueous solutions.20–23 Photosynthesis of urea not only makes use of abundant and free solar energy to alleviate the excessive use of fossil fuels and save energy, but also facilitates efficient and economical CO2 fixation, which is of great significance to solve the problems of resource scarcity, environmental issues, global warming. In particular, directly coupling N2 and CO2 to synthesize urea by photocatalysis have attracted much attention in recent years.24–26 In the process of photocatalytic urea synthesis, excited semiconductors first generate electrons and holes that participate in reduction–oxidation reactions to form nitrogen and carbon-containing intermediates. This is followed by the critical step of C–N coupling, and finally, electron-driven hydrogenation through proton-coupled reduction ultimately yields urea. Photocatalytic synthesis of urea is a prospective and innovative green chemical method, which opens up a sustainable pathway for urea synthesis.

Although photocatalytic synthesis of urea has been intensively investigated and made some progress in recent years, reported photocatalytic urea synthesis rates remain unsatisfactory. Currently, there are several challenges for the photosynthesis of valuable urea, such as high thermodynamic stability of reaction substrates (e.g., nitrogen and carbon dioxide),27–30 undesirable adsorption capacity of substrates/intermediates/products on the surface of the photocatalysts,31 high charge-carrier recombination efficiencies,32 slow kinetic processes of multi-electron transfers,33,34 and competing reactions in parallel products.35,36 These factors collectively result in low solar-to-chemical energy conversion efficiency, severely hindering the development of photocatalytic urea synthesis. In order to improve the photosynthesis performance of urea, a comprehensive understanding is essential to enable rational design of photocatalysts and systematic optimization of the photochemical reaction system. Herein, we present a systematic review of research advances in photocatalytic urea synthesis. The fundamental principles of the photocatalytic urea formation process are first introduced, covering the detection methods of key intermediates, as well as experimental identification and quantification techniques for urea and other parallel products. Subsequently, the performance of both classical and emerging photocatalysts across various reaction systems is summarized, with emphasis on their structure–activity relationships. Finally, current challenges in this field are discussed, and future research directions are outlined (Fig. 1). It is hoped that this review could provide an overview of the current research achievements related to the photocatalytic urea synthesis, thus stimulating the research on the design and synthesis of photocatalysts and the development of more efficient and stable photocatalytic systems.


image file: d5ra08170j-f1.tif
Fig. 1 Core framework of this perspective for urea photosynthesis.

2 The basic principle details for photocatalytic urea synthesis

2.1. Fundamentals of photocatalytic urea synthesis

Photocatalytic urea synthesis is a green synthetic method that utilizes renewable solar power to convert nitrogen-containing small molecules (e.g., N2, NO3, or NH3) and carbon sources (e.g., CO2, CO or CH3OH) into urea.37–40 Its basic principles involve multiple complex steps, mainly including: (1) the absorption of light energy by the photocatalyst, (2) the generation and separation of photogenerated carriers, (3) the adsorption and activation of reactants, and (4) the formation of the urea by C–N coupling reactions.

The main stages in photocatalytic urea synthesis are shown in Fig. 2. When a photocatalyst is exposed to light with an energy greater than its bandgap width, electrons in the valence band are excited and transition to the conduction band, forming photogenerated electron (e)–hole (h+) pairs. Then, the separation and transfer of photogenerated electrons and holes. There are several situations regarding the separation and transfer of photogenerated carriers. One is that the photoexcited electrons and holes are transferred to the reduction and oxidation sites on the semiconductor surface, where they undergo reduction and oxidation reactions with the adsorbed substrate molecules respectively, as in process ⑤, ⑥; another is that the excited state carriers undergo recombination during the process of separation and transfer, that is, bulk phase recombination, as in process ③; the third is the electron–hole recombination that occurs on the semiconductor surface, namely surface recombination, such as process ④. It is worth noting that the recombination of carriers severely hinders the solar energy conversion efficiency. Therefore, neither bulk phase nor surface recombination is conducive to the progress of photocatalytic reactions. Furthermore, adsorption and activation of reactants: carbon species and nitrogen sources are adsorbed on the surface of the photocatalyst and activated by photogenerated carriers. Finally, the activated intermediates undergo a C–N coupling reaction to form urea precursors which then proceed a series of hydrogenation and electron transfer processes to ultimately produce urea.


image file: d5ra08170j-f2.tif
Fig. 2 Schematic diagram of photocatalytic urea production.
2.1.1. Activation of carbon species. In the photocatalytic urea synthesis reaction, the reduction of CO2 is a critical step in forming the carbon source of urea. It is generally believed that CO2 is to form *CO intermediate, which then reacts with nitrogen-containing intermediates to generate urea. The specific analysis is as follows: the adsorbed CO2 molecule couple with one proton and one electron to form the adsorbed *HCOO (eqn (1)) which then further are reduced into *CO intermediate by the proton-assisted single electron transfer process (eqn (2)). The generated *CO intermediates couple with nitrogen-containing intermediates (such as activated N2, *NH2, *NO etc.) to form a C–N bond, ultimately generating urea. For CH3OH as the carbon source system, more·CH2OH radicals are generated from CH3OH, which are subsequently oxidized to *CHO intermediates on Pt cluster/TiO2.40
 
*CO2 + H+ + e → *COOH (1)
 
*COOH + H+ + e → *CO + H2O (2)
2.1.2. Activation of N-containing sources. The N2 molecule is the most stable known diatomic molecule, possessing a very short N[triple bond, length as m-dash]N bond length of only 109.8 pm with a bond strength of 941 kJ mol−1.41 A large energy gap (10.82 eV) exists between the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) of N2, whilst the molecule has both a high ionization potential (15.85 eV) and a low electron affinity (−1.9 eV) making oxidation or reduction difficult.42 The efficient activation of N2 is a key step in urea synthesis. N2 activation proceeds via two principal pathways: reduction and oxidation–reduction. In reduction systems, such as Cu SA-TiO2 (ref. 43), Pt cluster/TiO2 (ref. 40) and Pd–CeO2 (ref. 44), N2 is adsorbed onto metallic sites like Cu, Pt and Pd through “σ-π*” interactions and subsequently hydrogenated to yield *N2H, *N2H2, *NH2 (eqn (3)–(5)). Reductive sites in photocatalysts like Ti3+-TiO2, CeO2−x and Ru–O4Ti1 facilitate multi-electron transfer to directly crack the N[triple bond, length as m-dash]N bond. For the path of oxidation–reduction, N2 is first oxidized to nitrogen oxides, which are then coupled with *CO and gradually hydrogenated for reduction (eqn (6)–(8)). System like Ni1CdS/WO3,45 the N2 was converted into NO species by *OH radicals generated from photogenerated holes over the WO3 component, meanwhile, the CO2 was transformed into *CO species over the Ni site by photogenerated electrons. The generated NO and *CO species were further coupled to form *OCNO intermediate, then gradually transformed into urea. Other systems, such as nitrate and nitrite, mainly underwent hydrogenation reaction (eqn (9)–(15)).

N2 reduction:

 
*N2 + H+ + e → *N2H (3)
 
*N2H + H+ + e → *N2H2 (4)
 
*N2H + 2H+ + 2e → 2*NH2 (5)

N2 oxidation–reduction:

 
2H2O + 4h+ → 2OH + 2H + (6)
 
N2 + 2OH → 2NO + 2H + (7)
 
NO + 4H+ + 4e → *NH2 + H2O (8)

NO3 reduction:

 
*NO3 + H+ + e → *NO2OH (9)
 
*NO2OH + H+ + e → *NO2 + H2O (10)
 
*NO2 + H+ + e → *NOOH (11)
 
*NOOH + H+ + e → *NO + H2O (12)
 
*NO + 2H+ + 2e → *N + H2O (13)
 
*N + H+ + e → *NH (14)
 
*NH + H+ + e → *NH2 (15)

2.1.3. C–N coupling. Typically, the urea synthesis from CO2 and nitrogen-containing species is regarded as a co-reduction process. In general, C–N proceeds via three principal pathways. (1) *NH2 and *CO couple to form the *NH2CO (eqn (16)), which is followed by coupling with excessive *NH2 to produce urea. (2) *NO couples with *CO forming * ONCONO (eqn (17)). Subsequent stepwise hydrogenation of these intermediates yields urea (eqn (18)). (3) Activated *N[triple bond, length as m-dash]N couples with *CO to form *NCON (eqn (19)), which then undergoes hydrogenation (eqn (20)). Additionally, system like Pt Cluster/TiO2,40 the crucial step of C–N coupling was initiated by the reaction between *NH─NH and *CHO intermediate. It was observed that Pt clusters undergo “σ–π*” acceptor–donor interactions with N2, resulting in a reduction of the N2 activation barrier and subsequent activation into *NH–NH intermediates. Furthermore, under the promotion of N2 molecules and Pt clusters, more ·CH2OH radicals were generated from CH3OH, which were subsequently oxidized to *CHO intermediates. Further analysis with DFT calculation demonstrated that *NH–NH and *CHO were important precursors in C–N coupling reactions.
 
*NH2 + *CO → *NH2CO (16)
 
*NO + *CO → *ONCONO (17)
 
*ONCONO + 8H+ + 8e → (NH2)2CO + 2H2O (18)
 
*N2 + *CO → *NCON (19)
 
*NCON + 4H+ + 4e → (NH2)2CO (20)

2.2. Identification of key intermediates

Various intermediates are formed on the surface of the photocatalyst during activation of CO2 and nitrogen source, such as *CO2, *COOH, *CO, *N2, *N2H2, *NH2, *NH, *NO2, *NO2OH, *NOOH, *NO, ect. Among them, the determination of C–N coupling intermediates such as *NH2CO, *OCNO and *NCONO are particularly significant, which ultimately leads to urea formation. Therefore, the determination of key intermediates in the photocatalytic synthesis of urea is of great scientific significance for revealing the reaction mechanism and understanding the catalytic process.

Combining in situ characterization techniques with theoretical calculations can capture key intermediates, reveal active sites and reaction pathways, thereby deeply understanding the reaction mechanism and optimizing the catalytic system. In situ FTIR (DRIFTS) is an important technique for studying key intermediates in the photocatalytic synthesis of urea. It can monitor the dynamic evolution of adsorbed species on the catalyst surface in real time and reveal the reaction path. For example, Sheng et al. used in situ diffuse reflection infrared Fourier transform spectroscopy (in situ DRIFT) to monitor the photosynthesis process of urea. Upon illumination, the emergence and growth of several intermediate bands were observed in both cases (Fig. 3a). The two vibrational bands located at 1332 and 1173 cm−1 were attributed to the wagging and deformation modes of NH2 species, respectively, suggesting the one electron oxidation of NH3. For C–N intermediates, the 1439 cm−1 band corresponded to the C–N stretching vibration, closely resembling the standard band for urea, while the vibrational band located at 2205 cm−1 was assigned to the stretching mode of O[double bond, length as m-dash]C–NH2, derived from the coupling of NH2 with CO.46 Zheng et al. utilized in situ DRIFT spectroscopy to track the evolution of the critical intermediates in reactions.45 As illustrated in Fig. 3b, an obvious broadening band at about 1594 cm−1 was observed, which could be attributed to the co-existence of the stretching mode of adsorbed NO species at 1590 cm−1 and the adsorbed CO species at 1610 cm−1. More importantly, they captured the *OCNO intermediates at 2120 cm−1, which was generated by C–N coupling from *CO and *NO. Theoretical calculations can predict the stability of intermediates, energy barriers and the reaction mechanism, complementing experimental data. By calculating the adsorption configurations and energies of possible intermediates (such as *CO, *NH2, *CONH2, ect), thermodynamically feasible paths were screened out. For instance, Pang et al. performed DFT calculations to explore the photocatalytic C–N coupling mechanism for urea production (Fig. 3c).37 Comparing the Gibbs free-energy (ΔG) profiles for CO2 reduction, revealed that pure Fe2O3 faced challenges in generating *CO because of its inherently high CO2 activation barrier (with a rate-determining step of 1.28 eV). In contrast, CuL-Fe2O3 exhibited a more moderate rate-determining step (0.73 eV), thereby facilitating the generation of *CO intermediates. In addition, the desorption barrier stabilized the *CO intermediates for subsequent C–N coupling reactions. Considering the molecular orbital matching between *CO and *N2, the formation of *N2CO intermediates was a spontaneous exothermic reaction (Fig. 3d). Coupled with mass spectrometry (MS) and nuclear magnetic resonance spectroscopy (NMR), isotope labeling enables precise tracking of reaction pathways. Zhang et al. detected *NCONO intermediate (2100 cm−1) on Ru–O4Ti1 sites using In situ ATR-FTIR spectroscopy coupled with 15N-NMR (Fig. 3e–g).47 DFT calculations revealed unique Ru nanostructure (Ru–O4Ti1), which effectively triggered the activation of inert N2 molecules, facilitated the formation of crucial *NN(OH) intermediates (Fig. 3h), lowered the energy barrier of the potential determining step (Fig. 3i), and served as an “electronic pump” for electron migration from nitrogen to TiO2 support during urea photosynthesis.


image file: d5ra08170j-f3.tif
Fig. 3 (a) In situ FT-IR spectra of the photocatalytic urea synthesis under anaerobic (80% CO and 20% Ar) or aerobic (80% CO and 20% CO) condition on P25-4 h. Reproduced with permission from ref. 46. Copyright 2025, Wiley-VCH. (b) In situ DRIFT of Ni1–CdS/WO3 during photocatalytic urea synthesis. Reproduced with permission from ref. 45. Copyright 2024, Wiley-VCH. (c) Gibbs free-energy profiles of the CO2 reduction process for Fe2O3 and CuL-Fe2O3. (d) Gibbs free-energy profiles and the geometric structure diagram of intermediates evolution during the urea synthesis process for CuL-Fe2O3 by alternating hydrogenation pathway. Reproduced with permission from ref. 37. Copyright 2025, Elsevier. (e) In situ ATR-FTIR spectroscopy measurements of Ru–TiO2 over time during the coupling of N2 and CO2. The enlargement at 2100 cm−1 is shown in blue on the right. (f) 15N-NMR spectra of 15NO3 (standard) and test product from NOR using 15N2 as feeding gas. (g) 1H-NMR spectra of CO (15NH2)2 (standard) and test product from urea using 15N2 as feeding gas. (h) Calculated free energy diagram for possible NOR mechanism on TiO2 and Ru–TiO2 (blue line: TiO2; red line: Ru–TiO2). (i) Reaction barrier diagrams from *N2 to *NNOH on TiO2 and Ru–TiO2 (black line: TiO2; red line: Ru–TiO2). Reproduced with permission from ref. 48. Copyright 2024, Wiley-VCH.

2.3. Product identification and quantification

Photocatalytic urea synthesis is accompanied with the generation of various by-products, including CO, CH4, C2H4, HCOOH, CH3OH, C2H5OH, H2, NH3, CH3NH2, C2H5NH2, CH3CONH2, etc. Gas chromatography can be used to determine gas-phase products. For liquid-phase or soluble products, the identification and quantification could be achieved by nuclear magnetic resonance (NMR) spectroscopy, the diacetyl monoxime method or the urease decomposition method, high-performance liquid chromatography (HLPC) and high-performance liquid chromatography-mass spectrometry (LC-MS) spectroscopy.

For the typical identification and quantification of urea by 1HNMR spectroscopy, dimethyl sulfoxide-d6 (DMSO-d6) (99.8 atom % D with 0.03% [v/v] trimethylsilane) and disodium maleate aqueous solution as internal standards are added into reaction solution without any preliminary treatment. Subsequently, the mixed solution is transferred to a NMR tube and analyzed. However, the signal intensity of active hydrogen will be influenced by pH value and solution concentration.

Regarding the diacetyl monoxime method, typically, 5 g of diacetylmonoxime (DAMO) and 100 mg of thiosemicarbazide (TSC) are dissolved in distilled water to prepare the DAMO-TSC solution with 1000 mL in total volume. Subsequently, the preparation of acid-ferric solution is mainly to add 100 mL concentrated phosphoric acid and 300 mL concentrated sulfuric acid to 600 mL deionized water, and then dissolve 100 mg FeCl3 in the above solution. For the quantification of urea, 1 mL of DAMO-TSC solution and 2 mL of acid-ferric solution were added to 1 mL of sample solution. After vigorous mixing, the mixed solution was heated to 100 °C and maintained at this temperature for 15 minutes. When the solution is cooled to 25 °C, UV-vis absorption spectra were collected at a wavelength of 525 nm. The concentration–absorbance curve is calibrated using a standard urea solution with different concentrations. This method requires strict control of reaction conditions (temperature, time), and its accuracy can be affected by nitrates, ammonia, etc.

Additionally, the quantification of urea is achieved by HPLC (e.g. Agilent 1220 Infinity II) spectroscopy. HPLC was performed on a Luna 5 µm NH2 column (250 mm × 4.6 mm). The corresponding mobile phase, flow rate, and detected wavelength is needed. Measurement of urea in the reaction product: firstly, the product was rinsed repeatedly with ultrapure water to ensure that it was adequately collected, then pumped and filtered and transferred to a volumetric flask and diluted to a volume of 50 mL. 5 mL of it was concentrated to 1 mL for the assay. The advantages of the HPLC method are high resolution, good repeatability, and less impurity interference. Compared with HLPC, LC-MS has higher sensitivity and is suitable for trace analysis, and reaction mechanism research in photocatalytic synthesis of urea, but the instrument and maintenance costs are relatively high.

Overall, it is necessary to balance the detection target (urea concentration, matrix complexity), equipment availability and data requirements when choosing specific method. For mechanism research, NMR (isotope tracking) or LC-MS (product identification) are preferred. The HPLC and DAMO-TSC methods are recommended for routine testing inclinical/environmental applications. For trace detection, the HPLC and LC-MS methods offer ultra-high sensitivity for detecting urea.

3 The progress of urea photosynthesis

3.1. Urea synthesis from CO2 and NO3

In recent years, significant progress has been made in urea synthesis from inexpensive carbon and nitrogen sources. In particular, urea synthesis originates from CO2 and NO3 from industrial emission or agricultural fertilizer.4,15,48

Photochemical synthesis of urea was first reported by Yoneyama in 1998.49 Inspired by the high activities for reduction of CO2 to methanol and the favourable reaction rate for the reduction of NO3 to NH4+ of size-quantized TiO2 semiconductor nanocrystals immobilized in polyvinylpyrrolidinone gel film (Q-TiO2/PVPD), the group explored the simultaneous reduction of CO2 and NO3 (from LiNO3) to form urea using Q-TiO2/PVPD in the propylene carbonate solution containing isopropanol as a hole scavenger. Apart from the target urea, the by-products formed included methanol, ammonia, hydrogen, Ti3+ and acetone. Except for acetone, which was the only observed oxidation product derived from the oxidation of isopropyl alcohol, the other five products are reduction products. The sum of the quantum efficiencies of the reduction products (15.4) was very close to that of the oxidation products (15.9). Besides, NH2OH and NO were also used as nitrogen sources instead of NO3, and HCOOH and CO gas were used instead of CO2 as carbon sources. All cases tested in the study gave urea, but the reaction rate and distribution of products were largely influenced by the combination of carbon and nitrogen sources (CO2–NO3, CO–NO3, HCOOH–NO3, CO2–NH2OH, and CO2–NO) used. It should be noted that the reduction of NO3 was hypothesized as the rate-determining step in the urea photosynthesis reaction, as evidenced by the significant increase in urea yield when the nitrogen source is switched to NO or NH2OH, both of which are intermediates in the reduction reaction of NO3.

Subsequently, they studied that photocatalytic reduction of CO2 in the presence of nitrate ions using TiO2 nanocrystals embedded in SiO2 matrices.50 It was found that the urea production was influenced by the kind of solvents, including ethylene glycol monoethyl ether, acetonitrile, sulfolane, and water. From the results, the highest polarity solvent, which was water, yielded the highest amounts of urea. It was explained that these results may be due to the different degrees of dissociation of LiNO3 in solvents. With the increase of solvent polarity, the dissociation degree of LiNO3 increased, and the amount of NO3 available for the reduction reaction on TiO2 also increased. Aligned with their previous report, the primary conclusion of this study was that the reduction of NO3 was the rate-determining step in the photocatalytic production of urea.

In 2005, the photosynthesis of urea from CO2 and NO3 was reported by Shchukin et al. on TiO2/Cu particles encapsulated inside poly(styrene sulfonate)/poly(allylamine hydrochloride) capsules of different size (2.2, 4.2, and 8.1 µm) in aqueous solution.51 Poly(vinylalcohol) was employed as electron donor to facilitate the photosynthetic process. Fig. 4a showed an assembly scheme for spherical microreactor. The highest yield of urea photosynthesis (1.7 mM) was achieved for Cu-modified TiO2 nanoparticles encapsulated inside 2.2 µm poly(styrene sulfonate)/poly(allylamine hydrochloride) capsules (Fig. 4b). The decreasing the size of the confined microvolume of polyelectrolyte capsules accelerated the NO3 photoreduction, which was considered to be a limiting stage of the urea photosynthesis, and correspondingly improving the efficiency of urea production.


image file: d5ra08170j-f4.tif
Fig. 4 (a) Schematic illustration of the assembly of a photocatalytic microreactor. (b) Time course of the photoproduction of urea in CO2-saturated TiO2/Cu suspension (1) and inside TiO2/Cu-loaded polyelectrolyte capsules of 8.1 (2), 4.2 (3), and 2.2 µm (4) diameter. Reproduced with permission from ref. 51. Copyright 2005, American Chemical Society. (c) Initial rate of urea and formate photoproduction in CO2-saturated TiO2 suspension, suspension of Cu-modified TiO2, PFD-in-water emulsion stabilized with TiO2, and PFD-in-water emulsion stabilized with Cu-modified TiO2. Aquatic phase initially contained 1 M NaNO3 and 1 M isopropanol, pH 5.5. Reproduced with permission from ref. 37. Copyright 2005, Elsevier.

The same year, Ustinovich et al. reported that urea was synthesized by the photoinduced reduction of CO2 in the presence of NO3, using titania-stabilized perfluorodecalin-in-water (PFD:TiO2) emulsions and isopropanol as the hole scavenging reagent.52 The production of urea was obtained by irradiation PFD:TiO2 emulsion and TiO2 suspension, evidenced that in the case of emulsion the efficiency of urea photoproduction dramatically increased (Fig. 4c). Cu-modified nanodispersed TiO2 emulsion also studied the simultaneous photoreduction of CO2 and NO3, the urea photoproduction rate reached its maximum at copper loading of ca.3wt% and was higher than that of PFD:TiO2 emulsion (Fig. 4c). The increase of efficiency and selectivity of urea photoproduction observed for titania-stabilized perfluorocarbon-in-water emulsion can be attributed to high concentration of CO2 in the oleic phase contacting with photocatalyst and favourable conditions for stabilizing of the reaction intermediates to form C–N bonds in the case of two-phase reaction medium.

Despite the inspiring findings, early studies lacked discussion of the reaction mechanism. It was not until 2011 that Srinivas et al. first proposed a reaction pathway in his study on the photocatalytic synthesis of urea (Fig. 5a). Firstly, they investigated influence of hole scavengers (isopropanol and oxalic acid) on nitrate reduction for urea formation over TiO2 catalyst.12 The carbonaceous hole scavengers not only served as electron donors but also underwent further oxidation to produce CO2, thus serving as a carbon source, whereas the NO3 ion was used as the nitrogen source in their work (Fig. 5b). The results showed that the urea yield slightly increases under the isopropanol as the hole scavenger because the formation of ammonium oxalate, which was not desired under the oxalic acid as the hole scavenger, was avoided. To improve the adsorption of nitrate and to prevent the oxidation of ammonia, which was formed during nitrate reduction, the immobilization of TiO2 on zeolite was envisaged. The high activity of the TiO2 (10 wt%) supported over zeolite sample can be ascribed to the high dispersion of TiO2 on the surface, strong adsorption of substrates and also because of lower recombination of electron–hole pairs generated.


image file: d5ra08170j-f5.tif
Fig. 5 (a) Speculative mechanism for urea formation. (b) Schematic representation and stabilization of intermediate species during nitrate reduction and prevention of ammonia oxidation over TiO2-zeolite. Reproduced with permission from ref. 12. Copyright 2011, The American Society of Photobiology.

Huang et al. presented the novel Cs2CuBr4/TiOx-Ar (CCBT-Ar) photocatalyst for urea generation through the simultaneous reduction of CO2 and NO3.53 The synthesis of the designed CCBT-Ar catalysts was accomplished through a facile two-step method (as depicted in Fig. 6a). The X-ray diffraction (XRD) patterns showed that CCB was confined with amorphous TiOx (Fig. 6b). The scanning electron microscopy (SEM) of the CCBT-Ar adopted a spindle-like shape (Fig. 6c). The research findings demonstrated that the introduction of oxygen vacancies (Ov) in TiOx played a crucial role in reactant adsorption and the promotion of rate-determining steps in urea generation. However, the presence of Ov also led to significant carrier trapping effects, reducing the overall reaction activity. The well-designed CCBT-Ar catalyst exhibits outstanding photocatalytic urea generation activity, which attributes to its unique structural synergy. The TiOx nanocrystals were interconnected by in situ grown carbon nanosheets (Fig. 6d), allowing for efficient electron extraction from TiOx and acting as an electron reservoir, effectively suppressing Ov-induced carrier recombination. As electrons in TiOx were consumed, those reserved in the carbon nanosheets were replenished, ensuring the sustainability of the reaction.


image file: d5ra08170j-f6.tif
Fig. 6 (a) Schematic illustration of the synthesis procedure of CCBT-ar/air catalysts. (b) XRD patterns of CCBT-Ar, CCB, and TiOx-Ar. (c) SEM image of CCBT-Ar. (d) High-resolution TEM image of CCBT-Ar (with a 5 nm bar size inset). Reproduced with permission from ref. 53. Copyright 2025, Elsevier.

Li and coworkers engineered a TiO2 nanoparticle modified Cu nanorod photocatalyst (TiO2@Cu) for simultaneously promoting the NO3 reduction and CO2 reduction reaction in the photocatalytic synthesis of urea (Fig. 7a).54 The TiO2 nanoparticles were uniformly covered onto the surface of the Cu nanorod via a impregnation-reduction method (Fig. 7b), and the well-integrated core–shell TiO2@Cu showed excellent efficiency in photocatalytic urea synthesis, reaching up to 72.8 µmol g−1 h−1 of urea yield (Fig. 7c). The remarkable photoactivity was attributed to the unique Ti–O–Cu bond in heterojunction interface of TiO2@Cu (Fig. 7d–f), and Ti–O–Cu bond provided a favorable electron transfer pathway from TiO2 to Cu, which accelerated the transfer of photogenerated charge and reduced the recombination of hole and electron. Meanwhile, the introduction of Cu altered the energy band structure of TiO2, resulting in a smaller band gap and further improving the utilization of light. The density functional theory calculation (Fig. 7g) indicated that the energy barrier of the C–N coupling reaction in Ti–O–Cu site of TiO2@Cu (−3.22 eV) was much lower than individual Cu site (1.21 eV).


image file: d5ra08170j-f7.tif
Fig. 7 (a) Schematic diagram of photocatalytic co-reduction of nitrate ions and carbon dioxide for urea synthesis over TiO2@Cu nanorod. (b) Structure characterization. (SEM image of Cu, the SEM image of TC2, TEM image of TC2, HRTEM image of TC2, SAED image of TC2; the elemental mapping of TC2). (c) The yield of urea over TC2. (d) O 1s (e) Ti 2p and (f) Cu 2p for TC1, TC2, TC3, and TiO2 nanoparticle. (g) Free energy diagrams for the electrocatalytic C–N coupling of *CO and *NH2 on TiO2@Cu and Cu. Reproduced with permission from ref. 54. Copyright 2025, Tsinghua University Press.

It is hard to achieve the simultaneously catalytic reduction of two reactants CO2 and nitrate at a single active site. Therefore, constructing bimetallic active sites is a feasible method to simultaneously catalyze the reduction of CO2 and nitrate to C–N bonding.55 However, due to the long distance between the coupling reaction intermediates generated at the bimetallic sites in the alloy, it is difficult to achieve efficient coupling between CO2 and nitrate reduction intermediates. Hence, structuring bimetallic active site with short distance for the simultaneous catalytic co-reduction of CO2 and nitrate is conducive to the C–N coupling between reduction intermediate of CO2 and nitrate for urea synthesis. Zhao et al. engineered dual metal Cu and Ti active sites with a short distance of 2 Å by single atom Cu anchored on TiO2 toward photoelectrocatalytic urea synthesis from CO2 and nitrate.39 Cu and Ti dual active sites can efficiently catalyze the reduction of CO2 to *CO and reduction of nitrate to *NH2 intermediates, respectively. The relatively short distance of the Ti and Cu double site on SAC Cu–TiO2 was conducive to the coupling of the two reaction intermediates *CO and *NH2 to formation urea via C–N bonding by strong nucleophilic attack of *NH2. Density functional theory (DFT) calculations (Fig. 8) verified that compared with parallel competing reactions of *CO and *NH2 such as *CO hydrogenation, *NH2 hydrogenation and its dimerization, the coupling of *CO and *NH2 had a lower energy barrier on dual metal active sites with short distance by single atom Cu anchored on TiO2.


image file: d5ra08170j-f8.tif
Fig. 8 (a) Free energy of formation *CONH2 from *NH2 and *CO2, diagram of free energy changes and activation barriers of *CO and *NH2 coupled to CONH2 and other parallel CO2/NO3 reduction reaction on (b) SAC Cu–TiO2, (c) TiO2, (d) *CO(NH2)2 formation from CONH2 on TiO2 and SAC Cu–TiO2. Reproduced with permission from ref. 39. Copyright 2023, Elsevier.

The photoelectrochemical (PEC) system can be a viable solution for green synthesis of urea by combining the light absorbers and the catalysts into a fully integrated electrode. Thus, exploring and developing the PEC device is significant for green and efficient synthesis of urea in aqueous solution under mild conditions via solar energy. Wang et al. designed a hierarchical-structured Si-based photocathode to efficiently drive the solar-to-urea conversion by coupling the NO3and CO2, in which nanostructured n+p-Si serves as the light absorber decorated with TiO2 layer and NiFe diatomic cocatalysts on N-doped carbon nanosheets (Fig. 9a).56 The results demonstrated a remarkably high urea yield rate, faradaic efficiency (FE), and stable operation time of 81.1 mg h−1 cm−2, 24.2%, and 20 h at −1.0 V vs. RHE, respectively (Fig. 9b–d). The synergetic effect of NFDA, TiO2 layer, and n+p-Si enhances the charge-carrier dynamics of the photocathode. Duan et al. reported a photoelectrochemical method for urea synthesis by co-reduction of carbon dioxide and nitrate ion over a Cu2O photocathode,6 delivering urea formation rate of 29.71 ± 2.20 µmol g−1 h−1 and faradaic efficiency (FE) of 12.90 ± 1.15% at low external potential (0.017 V vs. Reversible hydrogen electrode) (Fig. 10a). The Cu2O exhibited cubic morphology and exposes (100) facet (Fig. 10b and c), with suitable conduction band (CB) position (at about 1.1 V vs. SHE) to co-reduce CO2 and NO3 (Fig. 10d–f). In situ Fourier transform infrared spectroscopy (Fig. 10h–j) and photo-assisted online differential electrochemical mass spectrometry (Fig. 10g) showed that CO* and image file: d5ra08170j-t1.tif species were the key intermediates for the subsequent C–N coupling. Density functional theory (DFT) calculations revealed that the first CN coupling during urea synthesis took places between CO* and image file: d5ra08170j-t2.tif, which were the rate determining step, and the second C–N coupling occurred between CONH* and image file: d5ra08170j-t3.tif. Zhao demonstrated an efficient GaN/Si photoelectrode for PEC urea synthesis from simultaneous NO3 and CO2 reduction reactions under solar light.57 The built-inpotential in n+−p Si and the inherent catalytic activity of GaN nanowires (NWs) led to the selective synthesis of urea at a low overpotential.


image file: d5ra08170j-f9.tif
Fig. 9 PEC urea synthesis of the Si-based photocathodes in a 0.1 M KHCO3+ 0.05 M KNO3 electrolyte. (a) Diagrammatic sketch of PEC urea synthesis under 1 sun illumination. (b) Urea yield rate (column diagrams) and FE (point plots) of NFDA/TiO2/Si (orange), NFDA/Si (blue), and TiO2/Si (green) at various potentials for 0.5 h. (c) Time dependence of urea yield (column diagrams) and FE (point plots) acquired from NFDA/TiO2/Si held at −1.0 V vs. RHE. (d) Jt curves of NFDA/TiO2/Si, NFDA/Si, and TiO2/Si at −1.0 V vs. RHE for 10 h. The Insets are the cross-sectional FESEM images of NFDA/TiO2/Si before and after 10-h PEC reactions. Reproduced with permission from ref. 56. Copyright 2024, PNAS.

image file: d5ra08170j-f10.tif
Fig. 10 (a) Urea formation rate and FE at different external potentials over Cu2O. (b) SEM image of Cu2O. (c) TEM image of Cu2O. (d) Diffuse reflectance ultraviolet-visible spectrum and the corresponding (αhν)2 versus photon energy plot of Cu2O. (e) VB XPS spectrum of Cu2O. (f) Mott–Schottky plots of Cu2O. (g) Photo-assisted DEMS measurements over Cu2O. Reaction conditions: (1) without external bias and CO2 in darkness; (2) with cathodic bias but without CO2 in darkness; (3) with cathodic bias but without CO2 under light irradiation (AM 1.5 G irradiation); (4) with cathodic bias and CO2 under light irradiation. In situ FTIR spectra in the range of (h) 1300–1800 cm−1, (i) 3150–3450 cm−1 and (j) 2050–2170 cm−1. Reproduced with permission from ref. 6. Copyright 2024, Wiley-VCH.

3.2. Urea synthesis from CO2 and N2

The conversion of N2 and CO2 into urea through a photocatalytic C–N coupling reaction under ambient conditions presents a favorable approach. On the one hand, N2 can serve as an abundant nitrogen source for urea photosynthesis, which accounts for 78% of the atmosphere. On the other hand, CO2, a main greenhouse gas leading to serious environmental concerns, can serve as a carbon source. Thus, the approach not only achieves effective energy conservation but also mitigates environmental concerns. Nevertheless, the photocatalytic urea production process still suffers from extremely serious challenges. Rationally designing photocatalysts that integrate the capture of inert gas molecules, bond cleavage and C–N coupling capabilities is a crucial prerequisite to promote the performance of urea photosynthesis.

In 2021, Maimaiti successfully established the photocatalytic synthesis of urea for the first time in the N2/CO2 system.58 They utilized oxygen vacancy-rich TiO2 loaded on carbon nanotubes with Fe cores (Ti3+–TiO2/Fe-CNTs) as the catalyst and achieved the coreduction of N2 and CO2 into urea in water without the addition of a hole scavenger. The authors identified Ti3+ sites and the adjacent oxygen vacancy serves as the active center for N2 and CO2, respectively. Adsorbed N2 and CO2 are further activated by photogenerated electrons, forming six-membered cyclic intermediates (H2NCONH2)2, which eventually evolved into urea (Fig. 11a).


image file: d5ra08170j-f11.tif
Fig. 11 (a) Mechanism of photocatalytic coreduction of N2/CO2 to form CO(NH2)2. Reproduced with permission from ref. 58. Copyright 2021, American Chemical Society. (b) Proposed mechanism of reversible and cooperative photocatalysis in Cu SA-TiO2. Reproduced with permission from ref. 43. Copyright 2022, Wiley-VCH. (c) Suggested mechanism for the photocatalytic synthesis of urea on 40% 2D-CdS@3D-BiOBr. Reproduced with permission from ref. 59. Copyright 2023, American Chemical Society. (d) Schematic diagram of the photocatalytic C–N coupling reaction in urea synthesis through the conversion of N2 and CO2. Reproduced with permission from ref. 44. Copyright 2023, The Royal Society of Chemistry.

It can be concluded from photocatalytic reactions involving multiple electrons (such as N2 fixation and CO2 reduction) that the efficient extraction of photogenerated electrons is an essential prerequisite for advanced photocatalysis. However, low availability of photogenerated electrons in intrinsic photocatalysts seriously hinders its further application. Zhang et al. reported a photoinduced strategy based on a TiO2 photocatalyst-immobilized reversible single-atom copper (denoted as Cu SA-TiO2) for the photocatalytic synthesis of urea using N2 and CO2 molecules in the presence of pure H2O (Fig. 11b).43 The introduction of reversible single-atom Cu in the designed sample imparted additional electron-rich sites for the photoactivation of reactants (N2, CO2, and H2O) and C–N coupling, together with the accelerated electron-transfer dynamics, ensuring the multi-electron supply for urea photosynthesis from N2, CO2, and H2O, thereby promoting urea photosynthesis.

Theoretically, integrating N2 reduction reaction semiconductors and CO2 reduction reaction semiconductors to construct composite photocatalytic materials is expected to enable the conversion of CO2 and N2 into urea under photocatalytic conditions. For example, Wang et al. prepared 2D-CdS@3D-BiOBr S-scheme heterostructures by self-assembly of BiOBr microspheres (3D-BiOBr) and CdS nanosheets (2D-CdS). It was proposed that Cd2+ in CdS and oxygen vacancies in BiOBr of 2D-CdS@3D-BiOBr hybrids facilitate the adsorption and activation of N2 and CO2, respectively, resulting in the formation of the *HNCONH intermediate.59 Subsequently, their group synthesized the CdS@Bi2WO6 heterojunction for photocatalytic CO2–N2–H2O to urea with visible light. It was suggested that urea was synthesized on the surface of CdS through two mechanisms: (I) the reaction between the activated intermediates of CO2 and N2 (mainly) and (II) the reaction of the in situ formed NH3 with the feed CO2 (Fig. 11c).

The weak adsorption and activation ability of inert gases (CO2 and N2) on photocatalysts has been the main challenge hindering the development of this technology.60–62 Niu designed a Pd-decorated CeO2 photocatalyst for the photoinduced coreduction of N2 and CO2 into urea,44 enabling spontaneous electron transfer at the palladium–ceria interface (Fig. 11d). The investigations further endorsed that the emerged space-charge region in the CeO2(111)/Pd(111) interface not only effectively facilitates the targeted capture and activation of inert CO2 and N2 but also stabilizes the formation of key intermediates (*NCON) (Fig. 12). Subsequently, they established mesoporous CeO2−x nanorods with adjustable oxygen vacancy concentration by heat treatment in Ar/H2 (90%:10%) atmosphere (Fig. 13a) and served as photocatalysts to convert both CO2 and N2 into urea under ambient conditions.63 By introducing oxygen vacancies to enhance the targeted adsorption and activation of N2 and CO2 (Fig. 13b), CeO2-500 (CeO2 nanorods heated treatment at 500 °C) revealed high photocatalytic activity toward the C–N coupling reaction for urea synthesis with a remarkable urea yield rate of 15.5 µg h−1 (Fig. 13c). In the most recent study, Meng synthesized a single-atom Ru and oxygen vacancies co-modified CeO2 (Ru1/CeO2–VO) by a photochemical strategy to achieve photocatalytic simultaneous reduction of CO2 and N2 for the synthesis of urea (Fig. 13d).64 Benefiting from the synergistic effect of Ce3+–VO and single atoms, Ru1/CeO2–VO nanosheets delivered the highest urea yield rate of 13.73 µmol g−1 h−1 (Fig. 13e). Strong electron metal support interactions between the Ru single atoms and CeO2 enabled effective separation of photoexcited carriers (Fig. 13f and g). Importantly, it was confirmed by experimental characterization and DFT calculations (Fig. 13h and i) that the vacancies can effectively adsorb CO2, and Ru single atoms can promote N2 adsorption and activation and contribute to hydrolysis dissociation, supplying protons for hydrogenation of active species.


image file: d5ra08170j-f12.tif
Fig. 12 (a) The adsorption energy of N2 at different reaction sites. (b) The adsorption energy of CO2 at different reaction sites. The differential charge density of *NCON at Pd(111)-CeO2 (111) interface. (c) Side view, (d) top view. (The blue color represents electron consumption and the yellow color represents electron accumulation). Reproduced with permission from ref. 44. Copyright 2023, The Royal Society of Chemistry.

image file: d5ra08170j-f13.tif
Fig. 13 (a) Schematic diagram of oxygen vacancy generation. (b) Adsorption energy of N2 and CO2 on CeO2 with oxygen vacancy. (c) The urea formation rates of CeO2-purchase (commercial), CeO2, CeO2-300, CeO2-500 and CeO2-700 with all optical spectrum as light source. Reproduced with permission from ref. 63. Copyright 2023, Wiley-VCH. (d) Mechanism of the photocatalytic urea synthesis based on synergistic effects of the Ru1/CeO2-VO. (e) The prepared materials and urea production efficiency under different conditions. (f) Photocurrent responses of CeO2, CeO2–VO, and Ru1/CeO2–VO. (g) PL spectra of CeO2, CeO2–VO, and Ru1/CeO2–VO. (h) N2 adsorption energy of different sites. (i) CO2 adsorption energy of different sites. Reproduced with permission from ref. 64. Copyright 2025, Wiley-VCH.

The use of metal elements as single active sites for the reduction of N2 or CO2 has been extensively studied and reported.65–68 However, urea synthesis requires the coupling reaction of the key intermediates *CO and *N to produce C–N bonds. Due to their high reactivity, these crucial reaction intermediates cannot exist in a stable form within the reaction system.42,69–71 Therefore, maintaining an optimal distance between the two reaction sites to enable the interaction of intermediates is an excellent strategy. Luo proposed a catalyst design strategy with dual active sites to meet the needs of urea synthesis (Fig. 14a). A series of composites (SiW12−XMoX@MIL101(Cr), X = 0, 3, 6, 9, 12) were obtained and applied for photocatalytic urea synthesis from N2 and CO2.72 The urea production rate of SiW6Mo6@MIL-101(Cr) reaches 1148 µg h−1 gcat−1 under the optimal experimental conditions (Fig. 14b). The performance from both experimental (Fig. 14c and d) and DFT calculation results (Table 1) indicated that the W site in SiW6Mo6 was assigned for the activation of nitrogen, whereas the Mo site was assigned for the activation of CO2. Zheng developed a multi-site photocatalyst, consisting of CeO2 nanorods decorated with Ru nanoparticles and Cu single atoms (Ru–Cu/CeO2), for the purpose of synthesizing urea at high yield.73 The incorporation of Ru and Cu sites was crucial not only to generate high-density photogenerated electrons, but also to facilitated N2 and CO2 adsorption and conversion. The in situ formed local nitrogen-rich area at Ru sites increased the encounter possibility with the carbon-containing species generated from Cu sites, substantially promoting C–N coupling (Fig. 14e). Lu reported a Cu single-atom-decorated porous Fe2O3 nanorod catalyst with a Cu–O–Fe configuration for the direct artificial photocatalytic synthesis of urea from N2 and CO2 in pure water.37 Because the d-orbitals of the Cu/Fe sites were close to the molecular orbitals of CO2/N2, the CuFe dual active sites can selectively adsorb and activate N2/CO2, thereby facilitating efficient C–N coupling (Fig. 14f and g).


image file: d5ra08170j-f14.tif
Fig. 14 (a) Schematic diagram of the design strategy for dual-active site catalysts in urea synthesis; (b) the urea production rate of composites in 100 ppm NaHCO3; (c) the NH4+ production rate of the composites in N2; (d) the CO production rates of the composites in 100 ppm NaHCO3. Reproduced with permission from ref. 72. Copyright 2024, The Royal Society of Chemistry. (e) The comparison between the designed Ru–Cu/CeO2 catalysts (above) and the reported multi-site catalysts (below) during urea synthesis. Reproduced with permission from ref. 73. Copyright 2025, Elsevier. (f) Difference charge-density stereograms of adsorbed N2 and CO2 on the Cu and Fe sites, respectively. Yellow and cyan represent electron accumulation and depletion, respectively. (g) Calculated projected density of states of the d-orbital of Cu and Fe in CuL-Fe2O3, and CO2/N2 molecular orbitals. Reproduced with permission from ref. 37. Copyright 2025, Elsevier.
Table 1 Preparation of the catalysts under different conditions. Reproduced with permission from ref. 72. Copyright 2024, The Royal Society of Chemistry
Species Etotal (ev) Emol−gas (ev) Esub-SiW6Mo6 (ev) Eads (ev)
Mo–CO2 471.519 −22.958 −448.034 −0.527
W–CO2 −471.303 −22.958 −448.034 −0.311
Mo–N2 −465.620 −17.096 −448.034 −0.490
W–N2 −465.968 −17.096 −448.034 −0.838


Photocatalysts with high reduction activity are necessary for overcoming the unfavorable energy barrier.42,74 It is well known that the Z-scheme photocatalyst can retain the high reduction capability of materials in the heterojunction.75–78 The high reduction activity favors boosting kinetics for co-reduction of N2 and CO2 and promoting C–N coupling reaction. Meanwhile, the Z-scheme photocatalyst has good charge separation ability, which can inhibit charge recombination and accelerate electron transfer.79–81 In this regard, the Z-scheme photocatalyst could be more reasonable for improving photocatalytic urea synthesis by co-reducing CO2 and N2. SrTiO3–FeS–CoWO4 Z-scheme photocatalyst was designed to promote urea synthesis by N2 and CO2 co-reduction in water (Fig. 15a).82 Its fast electron transfer overcomes the bottleneck of sluggish kinetics for C–N coupling reaction, and dual active sites for the adsorption and activation of N2 and CO2 enhanced the kinetics for urea synthesis, resulting in enhanced urea yield of 8054.2 µg gcat−1·h−1 on SrTiO3–FeS–CoWO4 (Fig. 15b and c).


image file: d5ra08170j-f15.tif
Fig. 15 (a) Energy band diagram of SrTiO3–FeS–CoWO4 Z-scheme heterojunctions. (b) Urea yields of SrTiO3–FeS–CoWO4 with different catalyst dosages; (c) urea yields of SrTiO3–FeS–CoWO4 with different gas-flow rates. Reproduced with permission from ref. 82. Copyright 2024, Wiley-VCH. (d) Comparison of the reaction conditions of different routes for urea synthesis. Reproduced with permission from ref. 48. Copyright 2024, Wiley-VCH. (e) Schematic illustration of photocatalytic urea synthesis mechanism over Ni1–CdS/WO3 catalyst. Reproduced with permission from ref. 45. Copyright 2024, Wiley-VCH.

The disparity in physical properties, structure, and catalytic kinetics between N2 and CO2 poses stringent demands for the design of key catalysts, reaction control, mass transfer, and other aspects. Considering the unique advantages of nitrogen oxidation activation and the bilateral redox of photocatalytic reactions, Ding et al. synthesized a Ru–TiO2 photocatalyst for urea photo-synthesis through a nitric acid-mediated pathway combining nitrogen oxidation and subsequent kinetically advantageous nitrate and CO2 co-reduction (Fig. 15d).48 The remarkable photo-activity was attributed to its unique oxygen vacancy-anchored Ru nanostructure (Ru–O4Ti1), which effectively activates inert N2 molecules to minish the disparity of orbital energy levels, facilitated the formation of crucial *NN(OH) intermediates, and served as an “electronic pump” to avoid electronegativity effect for facilitating electron transfer from nitrogen to TiO2 support for urea photosynthesis.

Typically, the urea synthesis from N2 and CO2 is regarded as a co-reduction process. Current works mainly focused on the photogenerated electrons, the role of photogenerated holes has been commonly ignored. If photogenerated holes could be employed in activating reactant, it would not only mitigate the competition between N2 and CO2 in photogenerated electrons, but also improve the utilization efficiency of photogenerated carriers toward urea product, thereby boosting the performance of photocatalytic urea synthesis. Zheng et al. constructed a redox heterojunction consisting of WO3 and Ni single-atom decorated CdS (Ni1–CdS/WO3),45 enabling photogenerated electrons and holes to participate in the conversion of CO2 and N2 respectively (Fig. 15e). Specifically, the N2 was activated by photogenerated holes and CO2 was converted by photogenerated electrons during photocatalytic urea synthesis, which mitigated the competition of photogenerated electrons between N2 and CO2, enhanced the utilization efficiency of photogenerated carriers for urea products, and ultimately promoted the synthesis of urea.

3.3. Urea synthesis from CO2 and NH3

Compared with the stable N[triple bond, length as m-dash]N bond in N2, the lone pair electrons in NH3 are naturally reactive centers, in favor of fast reaction dynamics.18,83 As a total result, photosynthesis of urea from NH3 and CO2 appears more promising for future practical industrial applications.83–87 Jiang et al. designed and synthesized a series of 3D N-heterocyclic covalent organic frameworks for urea photosynthesis from NH3 and CO2.38 Three isomorphic three-dimensional (3D) COFs with two fold interpenetrated ffc topology (namely3D-TPT-COF, 3D-PDDP-COF, and 3D-TBBD-COF) were functionalized by benzene, pyrazine, and tetrazine active cores, respectively, to modulate the catalytic micro-environment through the change in the number of heterocyclic N atoms on the active cores (Fig. 16a). The geometrical and electronic structural advantages endowed 3D-TBBD-COF with superior photocatalytic activity towards urea production with the yield of 523 µmol g−1 h−1, 40 and 4 times higher than that for 3D-TPT-COF and 3D-PDDP-COF, respectively (Fig. 16b), indicating the heterocyclic N microenvironment dependent catalytic performance for these COFs photocatalysts.
image file: d5ra08170j-f16.tif
Fig. 16 (a) Schematic synthesis of 3D-TPT-COF (green), 3D-TBBD-COF (red) and 3D-PDDP-COF (yellow); (b) urea formation rate using 3DTPT-COF, 3D-PDDP-COF, and 3D-TBBD-COF as photocatalyst. Reproduced with permission from ref. 38. Copyright 2025, Nature. (c) Dependence of urea yield and selectivity on the ratio of O2 in the atmosphere; EPR measurements of reactive radicals generated during photocatalysis on P25-4 h using DMPO as the trapping agent: (d) detection of ˙OH in pure aqueous solution under different atmospheric O2 ratios; (e) detection of O2˙ in methanol solution under different atmospheric O2 ratios; (f) detection of both ˙OH and ˙NH2 in 2 M ammonia aqueous solution under different atmospheric O2 ratios. (g) Comparisons of EPR signal obtained in 2 M ammonia under 20% O2 (red trace in (f)) with theoretically fitted EPR signals. (h) Performance comparison under the most optimized conditions (20% O2, 80% CO and pH = 9) and control systems lacking O2, CO, ammonia or light. Reproduced with permission from ref. 46. Copyright 2025, Wiley-VCH.

Besides, Sheng et al. presented a photocatalytic pathway for the selective urea synthesis through the oxidative coupling between CO and NH3.46 They uncovered that the O2 concentration played a crucial role in controlling both the urea production rate (Fig. 16c) and its selectivity by effectively controlling the generation of oxidative species (Fig. 16d–g), such as photogenerated holes (h+), superoxide radicals (O2˙) and hydroxyl radicals (˙OH). Using oxygen-deficient TiO2 under an air-level (20%) O2 atmosphere, a urea generation rate of 54.31 mg gcat−1 h−1 with 100% selectivity can be achieved (Fig. 16h). Mechanistic studies reveal that the process began with the oxidation of NH3 to ˙NH2 through oxidative radicals generated on TiO2, especially the oxygen-derived O2˙. This ˙NH2 radicals then coupled with CO to form urea.

3.4. Urea synthesis from other raw materials

The clean-energy-driven synthesis of urea from carbon- and nitrogen-containing small molecules has garnered significant interest but remained great challenges to achieve with high selectivity.

A Pt cluster-modified TiO2 (Pt cluster/TiO2) catalyst was designed through the impregnation reduction method to facilitate the photocatalytic synthesis of urea by promoting the simultaneous N2 reduction and CH3OH oxidation reactions (Fig. 17a).40 The prepared Pt cluster/TiO2 catalyst exhibited outstanding efficiency in urea synthesis, achieving a rate of 105.68 µmol g−1 h−1 with N selectivity of 97.29 ± 0.79% (Fig. 17b). Further analysis with density functional theory (DFT) calculation revealed that the “σ–π*” donor–acceptor interaction occurred between Pt clusters and N2 (Fig. 17c–e), efficiently reducing the N2 hydrogenation barrier. EPR experiments demonstrate that photogenerated electrons (e) and hole (h+) were synchronously consumed through N2 reduction and CH3OH oxidation (Fig. 17f–i), thereby accelerating urea synthesis. The crucial step of C–N coupling was initiated by the reaction between *NH–NH and *CHO intermediate, facilitated by the low energy barrier on Pt cluster/TiO2. For comparative analysis, the key performance parameters of photocatalysts employed in urea synthesis, including such metrics as urea yield, stability, and reaction conditions, are summarized in Table 2.


image file: d5ra08170j-f17.tif
Fig. 17 (a) Schematic diagram of the photocatalytic synergistic redox urea synthesis system; (b) blank control experiments of N2, CH3OH, and light for urea synthesis; (c) bader calculation of N2 on TiO2 and Pt cluster/TiO2; (d) a simplified schematic diagram of N2 bonding to a Pt center; (e) schematic diagram of N2 on the Pt cluster/TiO2 surface; (f) EPR results of the captured oxidative DMPO-CH2OH on the TiO2 and Pt cluster/TiO2; (g) EPR results of reductive TEMPO-e generation on the Pt cluster/TiO2 under N2 or Ar atmosphere; (h) EPR results of the captured oxidative DMPO-CH2OH on the Pt cluster/TiO2 under N2 and Ar atmosphere; (i) EPR results of reductive TEMPO-e generation on the Pt cluster/TiO2 with or without CH3OH during the photoredox reaction. Reproduced with permission from ref. 40. Copyright 2024, Wiley-VCH.
Table 2 Comparison of the urea production performance of reported photocatalysts. Reproduced with permission from ref. 23. Copyright 2025, Wiley-VCH
Photocatalysts Reactants Urea formation rate [µmol g−1 h−1] Light source AQY Cyclic stability
Cu SA-TiO2 N2, CO2, H2O 7.2 365 nm 10 h
Ni1–CdS/WO3 N2, CO2 7.8 385 nm 0.15% 10 h
2D-CdS@3D-BiOBr N2, CO2, H2O 15 Visible light 3.93% 2 h
TiO2/10 wt%–Fe2TiO5 NH3, CO2 3 UV 12 h
SrTiO3–FeS–CoWO4 N2, CO2 134.24 UV-vis 4 h
Pt cluster/TiO2 N2, CH3OH 105.68 9 h
Ru–TiO2 N2, CO2 24.95 380 nm 4.7% 3.75 h
Pd–CeO2 N2, CO2 9.2 UV-vis 15 h
CeO2−x N2, CO2 15.5 200–400 nm 3.93% 15 h
15%NiCoP–ZnIn2S4−x N2, CO2, H2O 19.6 420 nm 9.2% 9 h
Ti3+–TiO2 N2, CO2, H2O 177.53 UV-vis 12 h


4 Conclusions

Photocatalytic synthesis of urea via converting CO2 and N2/NH3/NO3 into high-value urea chemical under ambient conditions is a green and sustainable technology driven by clean and renewable energy. In this review, we systematically describe the basic principle details for photocatalytic urea synthesis, including fundamental mechanisms (light absorption by photocatalyst, separation and migration of the photogenerated charge carriers, surface redox reactions-activation of carbon and nitrogen reactants and C–N coupling), identification of key intermediates and product identification and quantification. We emphatically review experimental study, catalyst design and theoretical research progress on urea photosynthesis. Various systems, including CO2–N2, CO2–NO3 and CO2–NH3 systems, etc., have been systematically summarized. Urea has been successfully produced on photocatalyst-like metal oxides (TiO2, CeO2), single-atom (Cu SA-TiO2), MOF (Ce-BTC), inorganic composites (Ti3+–TiO2/Fe-CNTs) from the conversion of CO2 and types of nitrogen sources. To enhance the selectivity and efficiency of photosynthesis of urea, various photocatalysts have been designed and developed, including single-atom catalysts, bimetallic site catalysts, heterojunction catalysts and defect engineering, etc. The C–N coupling is a key step in urea synthesis, involving the activation and coupling of CO2 and nitrogen sources. The variation of catalysts and reaction conditions can lead to different reaction mechanisms. For instance, direct C–N coupling, where CO2 and N2 are directly coupled on the catalyst surface to form urea. In addition, indirect C–N coupling means that the nitrogen source is first reduced to intermediates (such as *NH2, *NO, etc.), and then coupled with *CO produced by the reduction of CO2 to form urea. In some cases, the photogenerated electrons and holes respectively activate CO2 and N2, promoting C–N coupling. Despite the diversity in raw materials, high-performance photocatalysts universally rely on a synergistic combination of three core aspects: (i) optimal adsorption and activation: the catalyst must possess active sites that can effectively adsorb and activate both N2 (or other N-sources) and CO2 (or other C-sources), often requiring dual or multiple active sites. (ii) Efficient charge separation and migration: a well-designed heterostructure or energy band alignment is crucial to spatially separate photogenerated electrons and holes and direct them to the respective reduction and oxidation sites. (iii) Facilitated C–N coupling: the local microenvironment and electronic structure of the active sites should be tuned to lower the energy barrier for the key C–N coupling step, which is often the rate-determining step. Although some exciting progress has been made, research in this area is still at the infant stage and requires further works. Firstly, the efficiency of C–N coupling is beyond satisfactory due to the difficulty in achieving the co-adsorption, activation and coupling processes of multiple reactants, such as the large difference in activation energy between CO2 and N2/NH3/NOx. Additionally, there is still controversy regarding the reaction mechanism, particularly concerning the activation of reactants, the formation of key intermediates, and the reaction pathway of the coupling step. Besides, the currently developed photocatalysts for urea photosynthesis mainly response to ultraviolet region. The inherently contradiction between more visible light absorption and the thermodynamic feasibility of urea photosynthesis is largely challenge due to a relatively high potential of the CO2RR to form *CO. An ideal future photocatalyst for urea synthesis should embody the following key properties: (i) broad-spectrum solar energy harvesting: the ability to utilize a wider range of the solar spectrum, including visible and even near-infrared light. (ii) Atomic-level precision in active sites: catalysts with well-defined, single-atom or cluster sites tailored for specific reactant adsorption and C–N coupling. (iii) Unprecedented selectivity: near 100% selectivity towards urea, effectively suppressing competing reactions (e.g., NH3 emission, H2 evolution). (iv) Exceptional long-term stability: robustness against photocorrosion, poisoning, and structural degradation over prolonged operation. Therefore, the design of multifunctional visible response photocatalysts, the development and application of advanced characterization techniques, and the exploration of novel reaction pathways require further research to address these challenges.

In summary, we hope that this review provides an overview of the current status and inspires greater interest in the development of alternative photocatalytic urea synthesis. Meanwhile, it is expected that through continuous innovation in photocatalyst design, optimization of reaction conditions, and in-depth understanding of reaction mechanisms, efficient photocatalytic urea synthesis can be achieved, making contributions to addressing energy crises and environmental issues.

Author contributions

Miss Peixia Li: writing original draft and conceptualization. Dr Zhidong Yang: supervision and writing review & editing.

Conflicts of interest

There are no conflicts of interest to declare.

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.

Acknowledgements

This project was financially supported by the Scientific Research Program Funded by Education Department of Shaanxi Provincial Government (No. 24JK0280), High-level Talents Foundation of Ankang University (No. 2023AYQDZR17).

References

  1. Y. Wang, D. Chen, C. Chen and S. Wang, Acc. Chem. Res., 2023, 57, 247–256 CrossRef PubMed .
  2. H. Wang, Z. Xin and Y. Li, Top. Curr. Chem., 2017, 375, 49 CrossRef .
  3. P. Xing, S. Wei, X. Chen, H. Luo, L. Dai and Y. Wang, Chem.–Eng. J., 2024, 494, 153135 CrossRef CAS .
  4. C. Lv, L. Zhong, H. Liu, Z. Fang, C. Yan, M. Chen, Y. Kong, C. Lee, D. Liu, S. Li, J. Liu, L. Song, G. Chen, Q. Yan and G. Yu, Nat. Sustain., 2021, 4, 868–876 CrossRef .
  5. D. Jiao, Y. Dong, X. Cui, Q. Cai, C. R. Cabrera, J. Zhao and Z. Chen, J. Mater. Chem. A, 2023, 11, 232–240 RSC .
  6. M. Li, Q. Shi, Z. Li, M. Xu, S. Yu, Y. Wang, S. M. Xu and H. Duan, Angew. Chem., Int. Ed., 2024, 63, e202406515 CrossRef CAS PubMed .
  7. P. Li, Q. Zhu, J. Liu, T. Wu, X. Song, Q. Meng, X. Kang, X. Sun and B. Han, Chem. Sci., 2024, 15, 3233–3239 RSC .
  8. Z. Zeng, C. Liao and L. Yu, Chin. Chem. Lett., 2024, 35, 109349 CrossRef CAS .
  9. C. Chen, X. Zhu, X. Wen, Y. Zhou, L. Zhou, H. Li, L. Tao, Q. Li, S. Du, T. Liu, D. Yan, C. Xie, Y. Zou, Y. Wang, R. Chen, J. Huo, Y. Li, J. Cheng, H. Su, X. Zhao, W. Cheng, Q. Liu, H. Lin, J. Luo, J. Chen, M. Dong, K. Cheng, C. Li and S. Wang, Nat. Chem., 2020, 12, 717–724 CrossRef CAS PubMed .
  10. D.-S. Huang, X.-F. Qiu, J.-R. Huang, M. Mao, L. Liu, Y. Han, Z.-H. Zhao, P.-Q. Liao and X.-M. Chen, Nat. Synth., 2024, 3, 1404–1413 CrossRef CAS .
  11. P. Xing, S. Wei, Y. Zhang, X. Chen, L. Dai and Y. Wang, ACS Appl. Mater. Interfaces, 2023, 15, 22101–22111 CrossRef CAS .
  12. B. Srinivas, V. D. Kumari, G. Sadanandam, C. Hymavathi, M. Subrahmanyam and B. R. De, Photochem. Photobiol., 2011, 88, 233–241 CrossRef .
  13. H. Cai, J. Ding, T. Hou, T. Wei, Q. Liu, J. Luo, L. Feng, W. Liu and X. Liu, Chem. Synth., 2024, 4, 100378 Search PubMed .
  14. N. Meng, Y. Huang, Y. Liu, Y. Yu and B. Zhang, Cell Rep. Phys. Sci., 2021, 2, 100378 CrossRef CAS .
  15. Y. Luo, K. Xie, P. Ou, C. Lavallais, T. Peng, Z. Chen, Z. Zhang, N. Wang, X.-Y. Li, I. Grigioni, B. Liu, D. Sinton, J. B. Dunn and E. H. Sargent, Nat. Catal., 2023, 6, 939–948 CrossRef CAS .
  16. Z. Mei, Y. Zhou, W. Lv, S. Tong, X. Yang, L. Chen and N. Zhang, ACS Sustain. Chem. Eng., 2022, 10, 12477–12496 CrossRef CAS .
  17. C. Mao, J. Byun, H. W. MacLeod, C. T. Maravelias and G. A. Ozin, Joule, 2024, 8, 1224–1238 CrossRef CAS .
  18. Y. Lou, H. Chen, L. Wang, S. Chen, Y. Song, Y. Ding, Z. Hao, C. He, D. Qiu, H. Li, J. Wang, D. Liu and X. Cui, ACS Sustain. Chem. Eng., 2025, 13, 151–164 CrossRef CAS .
  19. L. Kuang, Z. Chen, Y. Yan, F. Guo and W. Shi, Int. J. Hydrogen Energy, 2024, 87, 20–49 CrossRef CAS .
  20. J. Yuan, L. Hu, J. Huang, Y. Chen, S. Qiao and H. Xie, Appl. Catal., B, 2023, 339, 123146 CrossRef CAS .
  21. B. Bhattacharyya, R. M. Sarhan, Y. Lu and A. Taubert, ChemCatChem, 2024, 16, 202400638 CrossRef .
  22. H. Song, D. A. Chipoco Haro, P.-W. Huang, L. Barrera and M. C. Hatzell, Acc. Chem. Res., 2023, 56, 2944–2953 CrossRef CAS PubMed .
  23. Z. Lu, R. Chen, G. Liu, B. Xia, K. Fan, T. Liu, Y. Xia, S. Liu and B. You, Adv. Funct. Mater., 2025, 2500944 CrossRef CAS .
  24. Z. Zeng, Y. Chen, X. Zhu and L. Yu, Chin. Chem. Lett., 2023, 34, 107728 CrossRef CAS .
  25. J.-Y. Zeng, X.-S. Wang, X.-H. Liu, Q.-R. Li, J. Feng and X.-Z. Zhang, Natl. Sci. Rev., 2023, 10, nwad142 CrossRef CAS PubMed .
  26. C. Liu, F. Gong, Q. Zhou and Y. Xie, Energy Fuels, 2024, 38, 8951–8959 CrossRef CAS .
  27. S. Zhang, Y. Zhao, R. Shi, G. I. N. Waterhouse and T. Zhang, EnergyChem, 2019, 1, 100013 CrossRef .
  28. Z. Li, P. Zhou, M. Zhou, H. Jiang, H. Li, S. Liu, H. Zhang, S. Yang and Z. Zhang, Appl. Catal., B, 2023, 338, 122962 CrossRef CAS .
  29. Z. Yang, H. Zhang, J. Zhao, H. Shi, Y. Liu, H. Yang and P. Yang, ChemSusChem, 2022, 15, e202200260 CrossRef CAS PubMed .
  30. Z. Yang, Y. Zhang, H. Zhang, J. Zhao, H. Shi, M. Zhang, H. Yang, Z. Zheng and P. Yang, J. Catal., 2022, 409, 12–23 CrossRef CAS .
  31. S. N. Habisreutinger, L. Schmidt-Mende and J. K. Stolarczyk, Angew. Chem., Int. Ed., 2013, 52, 7372–7408 CrossRef CAS PubMed .
  32. S. Lian, M. S. Kodaimati, D. S. Dolzhnikov, R. Calzada and E. A. Weiss, J. Am. Chem. Soc., 2017, 139, 8931–8938 CrossRef CAS PubMed .
  33. P. Yang and Z. Yang, J. Energy Chem., 2020, 50, 365–377 CrossRef .
  34. Z. Guo, P. Zhou, L. Jiang, S. Liu, Y. Yang, Z. Li, P. Wu, Z. Zhang and H. Li, Adv. Mater., 2024, 36, 2311149 CrossRef CAS .
  35. W. Yang, R. Godin, H. Kasap, B. Moss, Y. Dong, S. A. J. Hillman, L. Steier, E. Reisner and J. R. Durrant, J. Am. Chem. Soc., 2019, 141, 11219–11229 CrossRef CAS PubMed .
  36. J. J. Leung, J. A. Vigil, J. Warnan, E. Edwardes Moore and E. Reisner, Angew. Chem., Int. Ed., 2019, 58, 7697–7701 CrossRef CAS .
  37. Y.-F. Mu, J.-L. Zhou, S.-X. Yuan, M.-R. Zhang, H. Pang, M. Zhang and T.-B. Lu, Chem Catal., 2025, 5, 101433 CAS .
  38. N. Li, J. Zhang, X. Xie, K. Wang, D. Qi, J. Liu, Y.-Q. Lan and J. Jiang, Nat. Commun., 2025, 16, 1106 CrossRef CAS PubMed .
  39. J. Zheng, S. Xu, J. Sun, J. Zhang, L. Sun, X. Pan, L. Li and G. Zhao, Appl. Catal., B, 2023, 338, 123056 CrossRef CAS .
  40. W. Yang, L. Xiao, W. Dai, S. Mou and F. Dong, Adv. Energy Mater., 2024, 14, 2303806 CrossRef CAS .
  41. B. Hu, B.-H. Wang, L. Chen, Z.-J. Bai, W. Zhou, J.-K. Guo, S. Shen, T.-L. Xie, C.-T. Au, L.-L. Jiang and S.-F. Yin, ACS Catal., 2022, 12, 11860–11869 CrossRef CAS .
  42. S. Lin, X. Zhang, L. Chen, Q. Zhang, L. Ma and J. Liu, Green Chem., 2022, 24, 9003–9026 RSC .
  43. D. Li, Y. Zhao, Y. Miao, C. Zhou, L. P. Zhang, L. Z. Wu and T. Zhang, Adv. Mater., 2022, 34, 2207793 CrossRef CAS PubMed .
  44. S. Yang, J. Deng, J. Chen, Q. Tan, T. Liu, K. Chen, D. Han, Y. Ma, M. Dai and L. Niu, Catal. Sci. Technol., 2023, 13, 1855–1865 RSC .
  45. Y. Zhang, Y. Sun, Q. Wang, Z. Zhuang, Z. Ma, L. Liu, G. Wang, D. Wang and X. Zheng, Angew. Chem., Int. Ed., 2024, 63, e202405637 CrossRef CAS .
  46. X. Huang, S. Xie, B. Sheng, B. Xiao, C. Chen, H. Sheng and J. Zhao, Angew. Chem., Int. Ed., 2025, 64, e202505630 CrossRef CAS PubMed .
  47. C. Shi, K. Xia, L. Zhang, M. Guo, X. Guan, C. Gu, X. Yang, Y. Wang, X. Liu and X. Ding, Angew. Chem., Int. Ed., 2024, 14, 2400201 CAS .
  48. M. Zhou, Y. Zhang, H. Li, Z. Li, S. Wang, X. Lu and S. Yang, Angew. Chem., Int. Ed., 2025, 64, e202414392 CrossRef CAS PubMed .
  49. H. Y. S. Kuwabata and H. Yoneyama, Langmuir, 1998, 14, 1899–1904 CrossRef .
  50. T. T. B. J. Liu and H. Yoneyama, J. Photochem. Photobiol., A, 1998, 115, 227–230 CrossRef .
  51. H. M. h. D. G. Shchukin, Langmuir, 2005, 21, 5582–5587 CrossRef PubMed .
  52. E. A. Ustinovich, D. G. Shchukin and D. V. Sviridov, J. Photochem. Photobiol., A, 2005, 175, 249–252 CrossRef CAS .
  53. H. Sun, Z. Lin, R. Tang, Y. Liang, S. Zou, X. Zhang, K. Chen, R. Zheng and J. Huang, Appl. Catal., B, 2025, 360, 124511 CrossRef CAS .
  54. R. Tan, S. Meng, P. Wang, C. Yang, J. Yao, H. Li, T. Zhang and Z. Li, Nano Res., 2025, 18, 94907647 CrossRef .
  55. Z. Ma, X. Xia, B. Song, R. Li, X. Wang and Y. Huang, ACS Appl. Mater. Interfaces, 2024, 16, 46323–46331 CrossRef CAS .
  56. Y. L. X. Zhang, C. Chena, J. Zheng, S. P. Jiang and S. Wang, Proc. Natl. Acad. Sci. U. S. A., 2024, 121, e2311326121 CrossRef .
  57. W. J. Dong, J. P. Menzel, Z. Ye, I. A. Navid, P. Zhou, K. R. Yang, V. S. Batista and Z. Mi, ACS Catal., 2024, 14, 2588–2596 CrossRef CAS .
  58. H. Maimaiti, B. Xu, J.-Y. Sun and L.-R. Feng, ACS Sustain. Chem. Eng., 2021, 9, 6991–7002 CrossRef CAS .
  59. Y. Wang, S. Wang, J. Gan, J. Shen, Z. Zhang, H. Zheng and X. Wang, ACS Sustain. Chem. Eng., 2023, 11, 1962–1973 CrossRef CAS .
  60. M. Yuan, J. Chen, Y. Bai, Z. Liu, J. Zhang, T. Zhao, Q. Wang, S. Li, H. He and G. Zhang, Angew. Chem., Int. Ed., 2021, 60, 10910–10918 CrossRef CAS .
  61. Y. Zhang, T. Hou, Q. Xu, Q. Wang, Y. Bai, S. Yang, D. Rao, L. Wu, H. Pan, J. Chen, G. Wang, J. Zhu, T. Yao and X. Zheng, Adv. Sci., 2021, 8, 2100302 CrossRef CAS .
  62. Y. He, L. Yin, N. Yuan and G. Zhang, Chem.–Eng. J., 2024, 481, 148754 CrossRef CAS .
  63. S. Yang, W. Zhang, G. Pan, J. Chen, J. Deng, K. Chen, X. Xie, D. Han, M. Dai and L. Niu, Angew. Chem., Int. Ed., 2023, 62, e202312076 CrossRef CAS .
  64. G. Ren, X. Chen, Z. Zhao, Z. Li and X. Meng, Adv. Funct. Mater., 2025, 2506296 CrossRef CAS .
  65. H. Liang, C. Ye, Y. Wu, Y. Li, R. Long, J. Xiong, W. Jiang and J. Di, Mater. Today, 2025, 86, 96–103 CrossRef CAS .
  66. H. Sun, X. Qin, Y. Zhang, Y. Xu, Y. Jiang, F. Pan and Q. Liu, Small, 2025, 21, 2503761 CrossRef CAS .
  67. F. Yang, H. Han, H. Duan, F. Fan, S. Chen, B. Yu Xia and Y. L. He, Adv. Energy Mater., 2025, 15, 2405726 CrossRef CAS .
  68. Y. Zhang, B. Xia, J. Ran, K. Davey and S. Z. Qiao, Adv. Energy Mater., 2020, 10, 1903879 CrossRef CAS .
  69. Y. Jiao, H. Li, Y. Jiao and S.-Z. Qiao, J. Am. Chem. Soc., 2023, 145, 15572–15580 CrossRef CAS .
  70. D. Anastasiadou and M. Costa Figueiredo, ACS Catal., 2024, 14, 5088–5097 CrossRef CAS .
  71. W.-Y. Lin, Z.-X. Chen, H. Xiong, H.-C. Li, Y.-S. Ho, C.-T. Hsieh, Q. Lu and M.-J. Cheng, ACS Catal., 2023, 13, 11697–11710 CrossRef CAS .
  72. S. Su, X. Li, W. Ding, Y. Cao, S. Yuan, Z. Liu, Y. Yang, Y. Ding and M. Luo, J. Mater. Chem. A, 2024, 12, 15300–15310 RSC .
  73. Q. Wang, Y. Wan, Q. Liu, Y. Zhang, Z. Ma, Z. Xu, P. Sun, G. Wang, H.-L. Jiang, W. Sun and X. Zheng, Sci. Bull., 2025, 70, 1118–1125 CrossRef CAS .
  74. J. Long, D. Luan, X. Fu, H. Li and J. Xiao, ACS Catal., 2024, 14, 14678–14687 CrossRef CAS .
  75. K. Qi, B. Cheng, J. Yu and W. Ho, Chin. J. Catal., 2017, 38, 1936–1955 CrossRef CAS .
  76. J. Li, H. Yuan, W. Zhang, B. Jin, Q. Feng, J. Huang and Z. Jiao, Carbon Energy, 2022, 4, 294–331 CrossRef CAS .
  77. Q. Xu, L. Zhang, J. Yu, S. Wageh, A. A. Al-Ghamdi and M. Jaroniec, Mater. Today, 2018, 21, 1042–1063 CrossRef CAS .
  78. X. Li, H. Sun, Y. Xie, Y. Liang, X. Gong, P. Qin, L. Jiang, J. Guo, C. Liu and Z. Wu, Coord. Chem. Rev., 2022, 467, 214596 CrossRef CAS .
  79. R. Sun, Z. Zhu, N. Tian, Y. Zhang and H. Huang, Angew. Chem., Int. Ed., 2024, 63, e202408862 CrossRef CAS .
  80. Z. Li, J. Xiong, Y. Huang, Y. Huang, G. I. N. Waterhouse, Z. Wang, Y. Mao, Z. Liang and X. Luo, Chem.–Eng. J., 2024, 486, 150304 CrossRef CAS .
  81. X. Fu, H. Huang, G. Tang, J. Zhang, J. Sheng and H. Tang, Chin. J. Struct. Chem., 2024, 43, 100214 CAS .
  82. M. I. Ahmad, Y. Liu, Y. Wang, P. Cao, H. Yu, H. Li, S. Chen and X. Quan, Angew. Chem., Int. Ed., 2024, 64, 167328 Search PubMed .
  83. X. Xiang, L. Guo, X. Wu, X. Ma and Y. Xia, Environ. Chem. Lett., 2012, 10, 295–300 CrossRef CAS .
  84. M.-A. Mohajer, P. Basuri, A. Evdokimov, G. David, D. Zindel, E. Miliordos and R. Signorell, Science, 2025, 388, 1426–1430 CrossRef CAS PubMed .
  85. I. P. Moura, A. C. Reis, A. E. Bresciani and R. M. B. Alves, Renew. Sustain. Energy Rev., 2021, 151, 111619 CrossRef CAS .
  86. J. Ding, R. Ye, Y. Fu, Y. He, Y. Wu, Y. Zhang, Q. Zhong, H. H. Kung and M. Fan, Nat. Commun., 2023, 14, 4586 CrossRef CAS PubMed .
  87. C. Zhao, Y. Jin, J. Yuan, Q. Hou, H. Li, X. Yan, H. Ou and G. Yang, J. Am. Chem. Soc., 2025, 147, 8871–8880 CrossRef CAS PubMed .

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