Skip to main content

Rational design of photocatalysts for ammonia production from water and nitrogen gas

Abstract

Photocatalytic N2 reduction has emerged as one of the most attractive routes to produce NH3 as a useful commodity for chemicals used in industries and as a carbon-free energy source. Recently, significant progress has been made in understanding, exploring, and designing efficient photocatalyst. In this review, we outline the important mechanistic and experimental procedures for photocatalytic NH3 production. In addition, we review effective strategies on development of photocatalysts. Finally, our analyses on the characteristics and modifications of photocatalysts have been summarized, based on which we discuss the possible future research directions, particularly on preparing more efficient catalysts. Overall, this review provides insights on improving photocatalytic NH3 production and designing solar-driven chemical conversions.

1 Introduction

Recently, there has been an increasing focus on ammonia (NH3) as not only a key commodity for chemicals widely used in industries [1, 2], but also as a liquid energy carrier [3] that enables transport and supply of hydrogen (H2) gas through cracking. Furthermore, NH3 is an alternative carbon-free energy source that can be utilized in energy-conversion devices, for example, direct NH3 fuel cells [4, 5]. Traditionally, NH3 production has mainly relied on the Haber–Bosch process, which is energy intensive and generates considerable volume of CO2 into the atmosphere because of these requirements, along with the needs for H2 and the extreme operation condition [6].

Photochemical NH3 production using N2 and water as a hydrogen source at ambient temperature and pressure can offer a promising alternative route that is less energy intensive and reduces CO2 emission, thereby significantly reducing environmental concerns [7]. NH3 production paired with water oxidation is a thermodynamically uphill reaction; therefore, external energy input is necessary [8]. Photocatalysts can utilize solar energy, which is supplied to Earth with sufficient solar power and abundance and facilitate NH3 production. Since TiO2 photocatalysts applied to N2 reduction, new photocatalysts and their modifications have been intensively suggested to improve photocatalytic NH3 production [9].

Major advances in metal oxides such as TiO2, WO3, and SrTiO3, bismuth oxyhalides, and polymeric carbon nitride photocatalysts have resulted in interests in the development of photocatalysts and their applications for NH3 production [10,11,12,13,14]. This review is composed of mainly two sections: (1) fundamentals for photocatalytic NH3 production, (2) strategies to develop photocatalysts. The first section focuses on the principles and mechanisms involved in the overall photocatalytic reaction. The second section contains recent progresses in the development of catalysts and their major advantages with regard to catalytic performance. The insights and discussions provided in this review will serve as useful resources for further development of photocatalysts and provide future research direction for photocatalytic NH3 production.

2 Fundamentals for photocatalytic NH3 production

2.1 Principle and mechanism

NH3 production by N2 reduction (N2(g) + 6H+  + 6e 2NH3(g), E0 = 0.092 V vs. SHE) paired with O2 production by water oxidation (2H2O(l)  O2(g) + 4H+  + 4e, E0 = 1.229 V vs. SHE) is a thermodynamically uphill reaction that requires a potential of at least 1.137 eV [15] per electron [16]. Moreover, three electrons are required to produce an NH3 molecule, and four holes are required to produce an O2 molecule according to eqns. (1) and (2):

$${\text{Absorption of light }}:{\text{photocatalyst}} + hv{ }\left( { > E_{{\text{g}}} } \right) \to e_{{{\text{CB}}}}^{ - } + {\text{h}}_{{{\text{VB}}}}^{ + }$$
$${\text{Nitrogen reduction }}\left( {{\text{NRR}}} \right) : {\text{N}}_{2} + 6{\text{H}}^{ + } + 6{\text{e}}_{{{\text{CB}}}}^{ - } \to 2{\text{NH}}_{3}$$
(1)
$${\text{Water oxidation }}\left( {{\text{OER}}} \right){ }:{ }2{\text{H}}_{2} {\text{O}} + 4{\text{h}}_{{{\text{VB}}}}^{ + } \to {\text{O}}_{2} + 4{\text{H}}^{ + }$$
(2)

A simplified schematic diagram for photocatalyic NH3 production and challenges in designing photocatalysts are summarized in Fig. 1. For overall NH3 production, photons with energies > 1.137 eV are required to generate photoexcited electron–hole pairs that carry out the desired redox reactions at the active sites on the surfaces of the photocatalysts. Thus, a photocatalyst must have an energy band gap, (Eg = ECBM-EVBM, where CBM and VBM represent conduction band minimum and valence band maximum), larger than the energy required for the overall NH3 production. In addition, a photocatalyst must have a suitable conduction and valence band alignment to drive two half-reactions using photoexcited electrons (e) and holes (h+) under solar light irradiation.

Fig. 1
figure 1

a The schematic illustration of photochemical nitrogen reduction reaction [15]. b The challenges and strategies for photocatalytic NH3 production in particulate photocatalytic (PC) ammonia production system

The major factors to consider for photocatalyst designs are (1) photon absorption followed by photoexcited carrier generation, (2) their migration and separation, and (3) their consumption by redox reactions on the surfaces of photocatalysts. The dominant photon energy spectrum obtained from the sunlight is in the range of 390–700 nm. Furthermore, because photocatalytic redox reactions require kinetic overpotentials, photocatalysts have a bandgap energy in the range of 1.6–2.4 eV. The generated photoexcited carriers can migrate to active sites on the surface where redox reactions carried out. However, since the kinetics of carrier recombination in the interiors of the photocatalysts is fast, considerable amounts of photoexcited carriers are recombined [17]. Moreover, photoexcited carriers can be trapped in the interior and surface of the photocatalyst, which causes photocorrosion. Thus, a photocatalyst design to improve electron–hole pair migration and separation is important for further development of photochemical NH3 production [18].

Most semiconductors that act as light absorbers in photocatalysts are inactive in the redox reactions with the photoexcited carriers; thus, the active sites on the surface of semiconductor are decorated with cocatalysts [19]. For example, cocatalysts, such as RuOx, CoOx, and Co-pi [20,21,22], combined on the surface of the semiconductor for facile water oxidation by providing active sites. Another critical point to note in NH3 production via photochemical N2 reduction is the facile competitive hydrogen evolution reaction by water reduction. NH3 production is kinetically much more complex than water reduction, although the thermodynamic requirements for those reactions are similar [23]. This indicates that photochemical H2 production may be predominant, resulting in inactive NH3 production. Thus, it is critical to modify the surface that provides active sites for the desired redox reaction using suppressing competitive reactions on the semiconductors.

It is also necessary to note the importance of rate balance for photocatalytic reactions on photocatalysts. Since electrons and holes are generated in pairs under illuminations, their consumption must be coupled as well. For instance, if oxidation reactions involving photogenerated holes are sluggish, the total NH3 production rate should be similar. In this case, scavengers, such as ethanol, methanol, and sulfites, can be employed as sacrificial reagents to facilitate oxidation reaction using photoexcited holes, thereby preventing from photoexcited carrier recombination [24]. However, since scavengers may interfere with counter reactions, here N2 reduction, or product evaluation methods, the target scavenger should be chosen carefully. It is discussed in detail in the next section.

2.2 Experimental process

The simplest photocatalytic NH3 production set-up is a particulate PC system where photocatalysts are dispersed in a medium, typically pure water, with N2 bubbling under simulated solar light illumination, as shown in Fig. 1b. NH3 and O2 are generated on a single photocatalyst. The produced NH3 is technically evaluated by the spectrophotometric method via the indophenol blue method or quantified using proton nuclear magnetic resonance spectroscopy (1H-NMR) with isotope labeled nitrogen 15 N as shown in Fig. 2.

Fig. 2
figure 2

a Spectrum and b standard curve of UV–Vis spectrophotometer with standard solutions from 0.5 to 5.0 µg·mL-1 in 0.05 M H2SO4. Reproduced with permission [25]. Copyright 2019, PNAS. c, d Interference impact on measurements of produced NH3 in solution with CH3OH, HCOOH and HCHO. Reproduced with permission [26]. Copyright 2019, Wiley–VCH. e 1H-NMR spectrums of 14NH4+(red) and 15NH4+(black), respectively. Reproduced with permission [27]. Copyright 2021, The Royal Society of Chemistry. f The peaks for 14NH4+ corresponding to its concentration. Reproduced with permission [8]. Copyright 2020, American Chemical Society

The indophenol blue method is based on the principle of the Berthelot reaction, where typically, salicylic acid, citrate, hypochlorite, and ferricyanide are used in an alkaline solution. In the presence of NH4+, the color of the yellowish indicating solution, which contains the reagents, turns to blue, corresponding to 655 nm. As shown in Fig. 2a and b, thus, the concentration of NH3 can be quantified by measuring the absorbance change at 655 nm using a UV–Vis spectrophotometer [28]. However, to quantify the concentration of NH3 generated by photocatalytic N2 reduction, multiple possible contaminants that interfere with the evaluation should be fully controlled [28]. For example, methanol, a widely used as a hole scavenger, and its derivatives cause incorrect color changes and significantly decrease the accuracy of NH3 quantification as shown in Fig. 2c and d. Another critical point for NH3, which is photocatalytically produced, quantification is to eliminate the parts from NH3 contaminants, unintentionally present introduced through catalyst, components in set-up [26] or even the air. In fact, NH3 present in human breath, sanitizers, and so can rapidly accumulate in water [29].

The most rigorous procedure to quantify photocatalytically generated NH3 from N2 reduction is to use isotope labeled 15N2 gas. The photocatalytic reactions are conducted in a simple particulate system with 15N2 bubbling. Since the splitting of 1H resonance in 15NH4+ and 14NH4+ resulting from the difference in the scalar interaction between N (15 N or 14 N) and H is differentiated, the 15NH4+ produced, which is dissolved in an aqueous reacting solution, can be quantified by 1H-NMR [30]. For instance, the 1H resonance in 15NH4+ can be split into two symmetric signals with a spacing of 73 Hz, whereas the resonance in 14NH4+ can be split into three symmetric signals with a spacing of 52 Hz as shown in Fig. 2e. Thus, false positive evaluations for the produced NH3 can be prevented effectively.

3 Strategies to develop photocatalysts

This section is divided into mainly two parts: (1) semiconductors comprising photocatalysts, which have been widely applied as a light absorber in Fig. 3a, and (2) strategies to modify the catalysts, which have been effectively used to further increase the photocatalytic NH3 production.

Fig. 3
figure 3

a The band alignments of widely used photocatalysts. b The Tauc plot (ahv)2 versus (hv) for the caluate the optical band gap energy Eg. c The color of Ni-x-TiO2 gradually deepens with the increase of the amount of Ni2+. d The N2-TCD spectrums of Ni-0.8-TiO2(red) and TiO2(black), respectively. e Bader charge density difference of Ni doped TiO2; Ti(silver), Ni(blue), and O(red). Reproduced with permission [45]. Copyright 2020, Americal Chemical Society

3.1 Semiconductors


Titanium(Ti)-based oxides Titanium dioxide, TiO2, possesses several advantages that make it attractive for use in photocatalytic NH3 production. TiO2 has a bandgap of ~ 3.2 eV, which absorbs UV light, and a CBM and VBM that offer sufficient overpotential for the photoexcited electrons to reduce N2 to NH3 and holes to oxidize water to O2, respectively. In addition, the inexpensive, environmentally benign, and stable nature of TiO2 is an additional advantage. Historically, the primary limitation of using TiO2 has been rapid photoexcited electron–hole recombination, which is similar to many other oxide-based photocatalysts. Strontium titanate, SrTiO3 [31], is another candidate showing two major benefits over TiO2 for use in the NH3 production. The CBM of SrTiO3 is slightly more negative than that of TiO2, thereby providing a stronger driving force to reduce N2 using the photoexcited electrons. In addition, intrinsic charge mobility in SrTiO3 is higher than that in TiO2, thereby suppressing photoexcited charge recombination [32]. However, despite this thermodynamic advantage, the photocatalytic NH3 production using TiO2 or SrTiO3 is limited because of ineffective light absorption resulting from their wide bandgaps and inactive redox reactions due to the limited active sites on their surface. Recently, many attempts have been made to address these limitations, which include doping, coupling with electrocatalysts (cocatalysts), surface defect engineering, and so on [33,34,35].


Bismuth oxyhalide (BiOX) BiOX (X = Cl, Br, and I) has a layered structure where bismuth and oxygen layers (BiOx) are connected with halide (X) layers alternatingly via Van der Waals forces [36]. Since the VBM is determined by O 2p and X \(\updelta\) p (\(\updelta\) = 3, 4, and 5, corresponding to X = Cl, Br, and I, respectively) and the CBM is mainly evaluated by Bi 6p, their band alignments and bandgap are varied depending on the element comprising the semiconductors [37]. For example, BiOCl, BiOBr, and BiOI have bandgap energies of 2.92, 2.65, and 1.75 eV, respectively [38], and are thermodynamically suitable for NH3 production, they enable the production of NH3 and O2 by the overall photocatalytic N2 reduction. However, despite these advantages, BiOX semiconductors have severe drawbacks when using them as photocatalysts, mainly due to the insufficient photogenerated charge separation in the bulk and surface. Recently, their fundamental limitations have been investigated, and strategies to improve their photocatalytic performances have been explored.


Graphitic carbon nitride (g-C3N4) g-C3N4 is a polymeric semiconductor with a bandgap energy of 2.7 eV and appropriate positions for the CBM and VBM for the NH3 and O2 production, respectively [39]. However, since rapid photoexcited carrier recombination in the bulk results in limited charge separation in the redox reactions and ineffective N2 adsorption on the surface, as an initial step for N2 reduction, multiple advancements to develop g-C3N4 have been suggested [40].

3.2 Strategies to enhance photocatalytic NH3 production

As described in the introduction section, photocatalysts comprising semiconductors can produce NH3 using N2 and water under illumination. However, the typical NH3 production rate and efficiency of intrinsic semiconductors using solar energy have been far lower than those expected from their bandgap energies. The major limiting factors include (1) limited light absorption because of the intrinsically large bandgaps, (2) significantly fast photoexcited electron–hole pairs recombination, and (3) slow redox reactions including N2 reduction and water oxidation. Recently, many strategies have been made to address one or more of these challenges, which include engineering defects by extrinsic doping or vacancy introducing, composites forming using multi-light absorbers, and coupling with electrocatalysts. This section will give an overview of each of these attempts and discuss how these strategies affected the photocatalytic NH3 production.

3.2.1 Defect engineering

Defect engineering has been widely suggested to control bulk and surface properties of catalysts, for example, by heteroatoms doping, vacancies formation, and so on. This defect engineering enhances photocatalytic NH3 production by changing the light absorption properties, improving charge separation efficiencies, or facilitating surface redox reactions, and it generally influences the multi-mechanistic steps of light harvesting simultaneously [41].


Heteroatom doping Heteroatom doping involves intentional introduction of impurity atoms into the lattices of materials to change their optical and electrical properties [42,43,44]. Depending on the species and degree (concentration) of dopants, changes, in the bandgap energy, band alignment, charge separation efficiency, and N2 adsorption ability are expected. However, since dopants can provide electron–hole recombination sites or prohibit charge transports, identifying types and concentrations of dopants appropriate for the photocatalytic process is important.

Zaicheng Sun and coworkers prepared a nickel-doped TiO2 (Ni-x-TiO2, where x = concentration of the Ni precursor controlled in the preparation step) photocatalyst using the sol–gel method [45] show Their optimized sample (Ni-0.8-TiO2) exhibited an NH3 production rate of 46.80 µmol·g−1·h−1 under simulated solar light irradiation, which is seven times higher than the rate using pure reference TiO2. The Ni-0.8-TiO2 sample extends the light absorption range to the visible light region with a bandgap energy of 2.92 eV, whereas pure TiO2 absorbs only UV light with a bandgap energy of 3.2 eV as shown in Fig. 3b. As Ni atoms having 2 + valence that replace the Ti site having 4 + valence, oxygen vacancies (Ov) are created naturally for charge neutrality within the host semiconductors, and they vary the band position of the CBM and VBM in TiO2, which results in a decrease in the bandgap energy. Besides changes in the optical properties, N2 adsorption, the initiation step for NH3 production from N2, is enhanced by Ni doping on TiO2, as investigated using a N2 temperature programmed desorption spectroscopy (N2-TPD) and a computational analysis.

Tierui Zhang and coworkers prepared a copper-doped TiO2 nanosheet (x%-TiO2, where x = molar ration of Cu/Ti controlled in the preparation step) photocatalyst using a hydrothermal method [46]. The optimized sample, 6%-TiO2, exhibited an NH3 production rate of 78.9 µmol·g−1·h−1, whereas pristine TiO2 nanosheets exhibited a rate of 0.34 µmol·g−1·h−1. As Cu heteroatoms replaced the sites of Ti in the TiO2 nanosheets, small amounts of Ti3+ valence states in Ti4+-dominant TiO2 and Ov sites are created, as observed using X-ray photoelectron spectroscopy and X-ray absorption fine structure spectroscopy. Furthermore, owing to the difference in size between the Cu2+ dopants and Ti4+ hosts, Cu-doped TiO2 has a compressive strain, which is evaluated by DFT calculations, and it causes changes in the distribution in the electron densities around O and Ti in the materials. The combination of non-stoichiometry and lattice strain in Cu-doped TiO2 narrows its bandgap energy, corresponding to the extended light absorption range of 600–800 nm, and electrons are accumulated around the sites of O atoms, thereby affecting the changes in the N2 adsorption affinity.

Jing Zhang and coworkers prepared a Fe-doped SrTiO3 (FexSr1-xTiO3, where x = the stoichiometric ratio of Fe precursor and Sr precursor controlled in preparation step) photocatalyst using the hydrothermal method followed by calcination [47]. The Fe0.1Sr0.9TiO3 sample exhibited the best NH3 production rate of 30.1 μmol·g−1·h−1, which is 3.2 times higher than that of pristine SrTiO3. By replacement of the Fe3+ ion, which has a smaller size, on the Sr2+ site in SrTiO3, the size of particles is reduced along with the corresponding increase in the surface area. Furthermore, the Fe0.1Sr0.9TiO3 photocatalyst exhibits significantly enhanced N2 chemisorption and activation ability, which Fe dopants are the major contributor to.


Vacancies Vacancy, one of defect formation strategies, leads to changes in the band structure and chemical adsorption nature on the surface [48]. This section gives an overview of intentionally induced anion vacancies in semiconductor photocatalysts and discusses how they affected the photocatalytic NH3 production.

Zhong Jin and coworkers compared the NH3 production activities using BiOBr semiconductors in the presence (Vo–BiOBr) or absence (BiOBr) of oxygen vacancies [49]. The Vo-BiOBr samples were prepared using the hydrothermal method with the addition of polyvinylpyrrolidone surfactants leading to in-situ generation of abundant vacancies, whereas pristine BiOBrs were synthesized without adding the surfactants. The Vo-BiOBr photocatalysts produced NH3 at a rate of 54.70 μmol·g−1·h−1 under UV–Vis irradiation, which is 10 times higher than the production rate using BiOBr. The major contribution of this significant improvement is enhanced N2 adsorption that initiates the intermediate formation for NH3 production. For example, the amount of adsorbed N2, estimated based on N2 adsorption isotherms, on Vo-BiOBr is considerably higher than that using BiOBr as shown in Fig. 4a. Furthermore, by introduction of the oxygen vacancies, the bandgap energy is reduced with the shifts in the CBM and VBM positions, which leads to increased light absorption as shown in Fig. 4b.

Fig. 4
figure 4

a The N2 adsorption isotherms, b The Tauc plot (ahv)2 versus (hv) and c The schematic illustration of pohtocatalytic NH3 synthesis process on Vo-BiOBr nanosheets and BiOBr nanoplates, respectively. Reproduced with permission [49]. Copyright 2018, American Chemical Society

Chuanyi Wang and coworkers prepared metal-free g-C3N4 semiconductors by thermal decomposition of melamine and induced nitrogen vacancies (V-g-C3N4) within the semiconductors by additional calcination under a N2 atmosphere [50]. The V-g-C3N4 samples produced NH3 at a rate of 1240 μmol·g−1 h−1 using pure N2 and water under visible light irradiation, whereas pristine g-C3N4 produced negligible amounts of NH3. Further, morphologies and bandgap energies were not significantly varied depending on the intentional creation of N vacancies; however, two factors changed considerably. Due to the presence of N vacancies on the semiconductor, the signal collected using photoluminescence spectra decreased, indicating the reduction in the carrier recombination and enhancement of their separation. Furthermore, the N vacancies enable selective adsorption and activate N2, and it was experimentally confirmed that negligible amount of NH3 produced on g-C3N4 selectively blocked the vacancies using Pd particles.

3.2.2 Coupling with multi-light absorbers

Coupling with multi-light absorbers to form heterojunctions has been widely investigated to improve light absorption capacity under single illumination and separation of photoexcited electron–hole carriers through their designed band alignments. This strategy for catalyst design is effective in improving the photocatalytic NH3 production.

Zhong Jin and coworkers prepared Bi2MoO6/Ov-BiOBr composites, where Ov denotes the oxygen vacancies using two-step methods, and synthesized Bi2MoO6 by solution-phase reflux process followed by the hydrothermal process to couple Ov-BiOBr on as-prepared Bi2MoO6 [51]. Despite minimal difference in light absorption before and after coupling, the Bi2MoO6/Ov-BiOBr composites showed a significantly enhanced NH3 production rate (90.7 μmol·g−1 h−1), which is almost 30 and 3 times higher than the rates evaluated using Bi2MoO6 and Ov-BiOBr, respectively as shown in Fig. 5a. The major role of coupling is the formation of cascaded band alignments at the interface between two semiconductors. The photoexcited holes are transported and oxidize the electron donor on Bi2MoO6 where the VBM is located relatively higher, whereas the photoexcited electrons are transported and reduce N2 on Ov-BiOBr where the CBM located relatively lower. Furthermore, oxygen vacancies on BiOBr show synergistic effects on the NH3 production by providing N2 adsorption and activation sites as shown in Fig. 5b.

Fig. 5
figure 5

a NH4+ concentrations of Bi2MoO6/Ov-BiOBr heterojunctions, Bi2MoO6 and BiOBr alone, respectively. b The scheme of proposed Bi2MoO6/BiOBr heterojunctions. Reproduced with permission [51]. Copyright 2019, The Royal Society of Chemistry. c The schematic reaction mechanism and d the artistic illustration of Au/TiO2-Ov photocatalyst. Reproduced with permission [52]. Copyright 2018, American Chemical Society

Jimmy Yu and coworkers prepared Ov-TiO2 using the hydrothermal method followed by loading Au nanoparticles. The Au nanoparticles generate hot electrons under visible light illumination [51], which is known as plasmonic phenomena. The hot electrons are injected into the CBM of TiO2 and then trapped in vacant sites where N2 is reduced into NH3. The generated holes remain on Au and oxidize the electron donors, methanol, in solution as shown in Fig. 5c and d. The optimized Au/TiO2-Ov samples produced NH3 with a rate of 78.6 μmol·g−1·h−1 under visible light irradiation, which is at least 98 and 35 times higher those using Au/TiO2 and TiO2-Ov photocatalysts, respectively.

3.2.3 Surface modification and reaction engineering

As discussed above, photocatalytic activities have been developed significantly by modifying photocatalysts, for example, defects engineering and composite formation. However, although photogenerated charge generation and separation are advanced by multiple possible strategies, if the surfaces of the semiconductors exhibit slow charge injection and are poorly catalytic for N2 reduction, the NH3 production rate may not be enhanced directly. The surface modification on the semiconductors using a robust electrocatalyst (cocatalyst) for N2 to NH3 conversion is the simplest way to improve the limiting factor at the interface between the semiconductors and liquid where reactants are present. Therefore, the modification of the semiconductor surface with various NH3 production catalysts, such as Ru, Cu, and Au, has been investigated [20, 34, 53].

In contrast, since the rates of reduction and oxidation reactions always correspond to one another, the sluggish photo-oxidation using holes may become a limiting factor for the photocatalytic NH3 production. To overcome the possible limitation, it has been widely suggested to load appropriate electrocatalysts (cocatalysts) for water oxidation on the semiconductor or provide a more facile hole acceptor (electron donors) than water, such as methanol or ethanol (Table 1).

Table 1 The summary of photocatalytic nitrogen reduction performances and its operating conditions: photocatalysts

4 Summary and outlook

In summary, we reviewed the mechanistic and experimental steps for photocatalytic NH3 production and examined various strategies to modify photocatalysts employed to enhance the production activities. This review clearly shows that although significant development of photocatalyts has been reported thus far, further advances for practical NH3 production are required. First, the efficiency for charge generation via solar light absorption should be enhanced. Considering that the majority of incident photons are in the range of visible light, development of absorbers, can harvest the incident solar light, is required by determining appropriate materials and engineering semiconductors.

The charge separation effect is also required for photocatalytic NH3 production through suppression of their recombination. For example, strategies for doping and composite formation will exhibit advances in the future and may increase the photon absorption efficiency simultaneously. In addition, morphology engineering [72,73,74,75,76] for photocatalysts will be promising and can be achieved by controlling the distances for carrier transportation and lifetime duration between the photo-induced generation and recombination. Furthermore, this may change facets and surface area, thereby providing active sites for the redox reactions using transported carriers. Moreover, if the facets are optimized with increasing surface area, morphology engineering can be significantly beneficial for NH3 production.

With improvements in photon absorption and charge separation, surface engineering will be able to provide active sites where N2 reduction using electrons and oxidation using holes is important. Thus, defect engineering, for example, formation of vacancies, cocatalyst coupling, and applying sacrificial reagents have been suggested. It is noteworthy that N2 diffusivity control close to photocatalysts make an important contribution to practical photocatalytic NH3 production. For instance, importance of three-phase interfaces, where a catalyst and two phases reactants such as water and active gas faced, has been demonstrated [7778]. This indicates that even if highly robust catalysts are prepared, if the supply of N2 gas and water is insufficient at the region close to the catalyst, the NH3 production is limited. Thus, environmental engineering, for example, via hydrophobicity control on the catalysts, will be significantly effective. The mechanism, experimental methods, discussion, insights, and suggestions contained in this review provide a good fundamental and foundation for further improvement of photocatalytic NH3 production.

Availability of data and materials

The datasets used and analysed during the current study are available from the corresponding references listed.

References

  1. J.W. Erisman, M.A. Sutton, J. Galloway, Z. Klimont, W. Winiwarter, How a century of ammonia synthesis changed the world. Nat. Geosci. 1(10), 636 (2008)

    Article  CAS  Google Scholar 

  2. V. Smil, Detonator of the population explosion. Nature 400(6743), 415 (1999)

    Article  CAS  Google Scholar 

  3. J.G. Chen, R.M. Crooks, L.C. Seefeldt et al., Beyond fossil fuel-driven nitrogen transformations. Science 360(6391), eaar6611 (2018)

    Article  CAS  Google Scholar 

  4. Y.J. Jang, K.S. Choi, Enabling electrochemical N2 reduction to NH3 in the low overpotential region using non-noble metal Bi electrodes via surface composition modification. J Mater Chem A 8(27), 13842 (2020)

    Article  CAS  Google Scholar 

  5. D.R. MacFarlane, P.V. Cherepanov, J. Choi, B.H.R. Suryanto, R.Y. Hodgetts, J.M. Bakker, F.M.F. Vallana, A.N. Simonov, A roadmap to the ammonia economy. Joule 4(6), 1186 (2020)

    Article  CAS  Google Scholar 

  6. J. John, D.K. Lee, U. Sim, Photocatalytic and electrocatalytic approaches towards atmospheric nitrogen reduction to ammonia under ambient conditions. Nano Converg. 6(1), 15 (2019)

    Article  Google Scholar 

  7. K. Ithisuphalap, H. Zhang, L. Guo, Q. Yang, H. Yang, G. Wu, Photocatalysis and photoelectrocatalysis methods of nitrogen reduction for sustainable ammonia synthesis. Small Methods 3(6), 1800352 (2019)

    Article  CAS  Google Scholar 

  8. Y.J. Jang, A.E. Lindberg, M.A. Lumley, K.S. Choi, Photoelectrochemical nitrogen reduction to ammonia on cupric and cuprous oxide photocathodes. ACS Energy Lett. 5(6), 1834 (2020)

    Article  CAS  Google Scholar 

  9. G.N. Schrauzer, T.D. Guth, Photolysis of water and photoreduction of nitrogen on titanium dioxide. J Am Chem. Soc. 99(22), 7189 (1977)

    Article  CAS  Google Scholar 

  10. G. Zhang, X. Yang, C. He, P. Zhang, H. Mi, Constructing a tunable defect structure in TiO2 for photocatalytic nitrogen fixation. J. Mater. Chem. A 8(1), 334 (2020)

    Article  CAS  Google Scholar 

  11. E. Endoh, J.K. Leland, A.J. Bard, Heterogeneous photoreduction of nitrogen to ammonia on tungsten oxide. J. Phys. Chem 90(23), 6223 (1986)

    Article  CAS  Google Scholar 

  12. Q.S. Li, K. Domen, S. Naito, T. Onishi, K. Tamaru, Photocatalytic synthesis and photodecomposition of ammonia over SrTiO3 and BaTiO3 based catalysts. Chem. Lett. 12(3), 321 (1983)

    Article  Google Scholar 

  13. Y. Zhao, S. Zhou, J. Zhao, Y. Du, S.X. Dou, Control of photocarrier separation and recombination at bismuth oxyhalide interface for nitrogen fixation. J. Phys. Chem. Lett. 11(21), 9304 (2020)

    Article  CAS  Google Scholar 

  14. L. Yu, Z. Mo, X. Zhu, J. Deng, F. Xu, Y. Song, Y. She, H. Li, and H. Xu, Construction of 2D/2D Z-scheme MnO2-x/g-C3N4 photocatalyst for efficient nitrogen fixation to ammonia, Green Energy Environ., (2020)

  15. H. Gal, G. Alan, F. Frank A. et al., The potential economic feasibility of direct electrochemical nitrogen reduction as a route to ammonia, ChemRxiv, (2019)

  16. M. Pourbaix, Atlas of electrochemical equilibria in aqueous solutions (National Association of Corrosion Engineers, Houston, Tex., 1974).

  17. L. Zhang, H.H. Mohamed, R. Dillert, D. Bahnemann, Kinetics and mechanisms of charge transfer processes in photocatalytic systems: a review. J. Photochem. Photobiol. C-Photochem. Rev. 13(4), 263 (2012)

    Article  CAS  Google Scholar 

  18. M.A. Lumley, A. Radmilovic, Y.J. Jang, A.E. Lindberg, K.S. Choi, Perspectives on the development of oxide-based photocathodes for solar fuel production. J. Am. Chem. Soc. 141(46), 18358 (2019)

    Article  CAS  Google Scholar 

  19. W. Zhao, J. Zhang, X. Zhu, M. Zhang, J. Tang, M. Tan, Y. Wang, Enhanced nitrogen photofixation on Fe-doped TiO2 with highly exposed (101) facets in the presence of ethanol as scavenger. Appl. Catal. B-Environ. 144, 468 (2014)

    Article  CAS  Google Scholar 

  20. S. Liu, Y. Wang, S. Wang et al., Photocatalytic fixation of nitrogen to ammonia by single Ru atom decorated TiO2 nanosheets. ACS Sustain. Chem. Eng. 7(7), 6813 (2019)

    Article  CAS  Google Scholar 

  21. Z. Yan, H. Wu, A. Han, X. Yu, P. Du, Noble metal-free cobalt oxide (CoOx) nanoparticles loaded on titanium dioxide/cadmium sulfide composite for enhanced photocatalytic hydrogen production from water. Int. J. Hydrog. Energy 39(25), 13353 (2014)

    Article  CAS  Google Scholar 

  22. X. Yue, S. Yi, R. Wang, Z. Zhang, S. Qiu, Cobalt phosphide modified titanium oxide nanophotocatalysts with significantly enhanced photocatalytic hydrogen evolution from water splitting. Small 13(14), 1603301 (2017)

    Article  CAS  Google Scholar 

  23. S. Chen, D. Liu, T. Peng, Fundamentals and recent progress of photocatalytic nitrogen-fixation reaction over semiconductors. Sol. RRL 5(2), 2000487 (2021)

    Article  CAS  Google Scholar 

  24. S. Zhang, Y. Zhao, R. Shi, G.I.N. Waterhouse, T. Zhang, Photocatalytic ammonia synthesis: recent progress and future. EnergyChem 1(2), 100013 (2019)

    Article  Google Scholar 

  25. P. Chen, N. Zhang, S. Wang et al., Interfacial engineering of cobalt sulfide/graphene hybrids for highly efficient ammonia electrosynthesis. Proc. Natl. Acad. Sci. U. S. A.116(14), 6635 (2019)

    Article  CAS  Google Scholar 

  26. Y. Zhao, R. Shi, X. Bian et al., Ammonia detection methods in photocatalytic and electrocatalytic experiments: how to improve the reliability of NH3 production rates? Adv. Sci. 6(8), 1802109 (2019)

    Article  CAS  Google Scholar 

  27. M. Cong, X. Chen, K. Xia, X. Ding, L. Zhang, Y. Jin, Y. Gao, L. Zhang, Selective nitrogen reduction to ammonia on iron porphyrin-based single-site metal–organic frameworks. J. Mater. Chem. A 9(8), 4673 (2021)

    Article  CAS  Google Scholar 

  28. S.Z. Andersen, V. Čolić, S. Yang et al., A rigorous electrochemical ammonia synthesis protocol with quantitative isotope measurements. Nature 570(7762), 504 (2019)

    Article  CAS  Google Scholar 

  29. L. Li, C. Tang, D. Yao, Y. Zheng, S.Z. Qiao, Electrochemical nitrogen reduction: identification and elimination of contamination in electrolyte. ACS Energy Lett. 4(9), 2111 (2019)

    Article  CAS  Google Scholar 

  30. A.C. Nielander, J.M. McEnaney, J.A. Schwalbe et al., A versatile method for ammonia detection in a range of relevant electrolytes via direct nuclear magnetic resonance techniques. ACS Catal. 9(7), 5797 (2019)

    Article  CAS  Google Scholar 

  31. F. Wu, Y. Yu, H. Yang et al., Simultaneous enhancement of charge separation and hole transportation in a TiO2–SrTiO3 core–shell nanowire photoelectrochemical system. Adv. Mater. 29(28), 1701432 (2017)

    Article  CAS  Google Scholar 

  32. B. Huang, Y. Liu, Q. Pang, X. Zhang, H. Wang, P.K. Shen, Boosting the photocatalytic activity of mesoporous SrTiO3 for nitrogen fixation through multiple defects and strain engineering. J. Mater. Chem. A 8(42), 22251 (2020)

    Article  CAS  Google Scholar 

  33. Z. Ying, S. Chen, S. Zhang, T. Peng, R. Li, Efficiently enhanced N2 photofixation performance of sea-urchin-like W18O49 microspheres with Mn-doping. Appl. Catal. B-Environ. 254, 351 (2019)

    Article  CAS  Google Scholar 

  34. P. Huang, W. Liu, Z. He et al., Single atom accelerates ammonia photosynthesis. Sci. China-Chem. 61(9), 1187 (2018)

    Article  CAS  Google Scholar 

  35. S. Hu, X. Chen, Q. Li, Y. Zhao, W. Mao, Effect of sulfur vacancies on the nitrogen photofixation performance of ternary metal sulfide photocatalysts. Catal. Sci. Technol. 6(15), 5884 (2016)

    Article  CAS  Google Scholar 

  36. D.S. Bhachu, S.J.A. Moniz, S. Sathasivam et al., Bismuth oxyhalides: synthesis, structure and photoelectrochemical activity. Chem. Sci. 7(8), 4832 (2016)

    Article  CAS  Google Scholar 

  37. J. Li, H. Li, G. Zhan, L. Zhang, Solar water splitting and nitrogen fixation with layered bismuth oxyhalides. Accounts Chem. Res. 50(1), 112 (2017)

    Article  CAS  Google Scholar 

  38. W.L. Huang, Electronic structures and optical properties of BiOX (X = F, Cl, Br, I) via DFT calculations. J. Comput. Chem. 30(12), 1882 (2009)

    Article  CAS  Google Scholar 

  39. S. Liu, S. Wang, Y. Jiang, Z. Zhao, G. Jiang, Z. Sun, Synthesis of Fe2O3 loaded porous g-C3N4 photocatalyst for photocatalytic reduction of dinitrogen to ammonia. Chem. Eng. J. 373, 572 (2019)

    Article  CAS  Google Scholar 

  40. R. Liu, Z. Chen, Y. Yao, Y. Li, W.A. Cheema, D. Wang, S. Zhu, Recent advancements in g-C3N4-based photocatalysts for photocatalytic CO2 reduction: a mini review. RSC Adv. 10(49), 29408 (2020)

    Article  Google Scholar 

  41. R. Shi, Y. Zhao, G.I.N. Waterhouse, S. Zhang, T. Zhang, Defect engineering in photocatalytic nitrogen fixation. ACS Catal. 9(11), 9739 (2019)

    Article  CAS  Google Scholar 

  42. Y.J. Jang, M.D. Bhatt, J. Lee, S.H. Choi, B.J. Lee, J.S. Lee, Metal-free artificial photosynthesis of carbon monoxide using N-doped ZnTe nanorod photocathode decorated with N-doped carbon electrocatalyst layer. Adv. Energy Mater. 8(20), 1702636 (2018)

    Article  CAS  Google Scholar 

  43. J.H. Kim, Y.J. Jang, S.H. Choi, B.J. Lee, J.H. Kim, Y.B. Park, C.M. Nam, H.G. Kim, J.S. Lee, A multitude of modifications strategy of ZnFe2O4 nanorod photoanodes for enhanced photoelectrochemical water splitting activity. J. Mater. Chem. A 6(26), 12693 (2018)

    Article  CAS  Google Scholar 

  44. Y.J. Jang, Y.B. Park, H.E. Kim, Y.H. Choi, S.H. Choi, J.S. Lee, Oxygen-intercalated CuFeO2 photocathode fabricated by hybrid microwave annealing for efficient solar hydrogen production. Chem. Mat. 28(17), 6054 (2016)

    Article  CAS  Google Scholar 

  45. J. Li, D. Wang, R. Guan, Y. Zhang, Z. Zhao, H. Zhai, Z. Sun, Vacancy-enabled mesoporous TiO2 modulated by nickel doping with enhanced photocatalytic nitrogen fixation performance. ACS Sustain. Chem. Eng. 8(49), 18258 (2020)

    Article  CAS  Google Scholar 

  46. Y. Zhao, Y. Zhao, R. Shi, B. Wang, G.I.N. Waterhouse, L.Z. Wu, C.H. Tung, T. Zhang, Tuning oxygen vacancies in ultrathin TiO2 nanosheets to boost photocatalytic nitrogen fixation up to 700 nm. Adv. Mater. 31(16), 1806482 (2019)

    Article  CAS  Google Scholar 

  47. Z. Ying, S. Chen, T. Peng, R. Li, J. Zhang, Fabrication of an Fe-doped SrTiO3 photocatalyst with enhanced dinitrogen photofixation performance. Eur. J. Inorg. Chem. 2019(16), 2182 (2019)

    Article  CAS  Google Scholar 

  48. M. Li, H. Huang, J. Low, C. Gao, R. Long, Y. Xiong, Recent progress on electrocatalyst and photocatalyst design for nitrogen reduction. Small Methods 3(6), 1800388 (2019)

    Article  CAS  Google Scholar 

  49. X. Xue, R. Chen, H. Chen et al., Oxygen vacancy engineering promoted photocatalytic ammonia synthesis on ultrathin two-dimensional bismuth oxybromide nanosheets. Nano Lett. 18(11), 7372 (2018)

    Article  CAS  Google Scholar 

  50. G. Dong, W. Ho, C. Wang, Selective photocatalytic N2 fixation dependent on g-C3N4 induced by nitrogen vacancies. J. Mater. Chem. A 3(46), 23435 (2015)

    Article  CAS  Google Scholar 

  51. X. Xue, R. Chen, C. Yan, Y. Hu, W. Zhang, S. Yang, L. Ma, G. Zhu, Z. Jin, Efficient photocatalytic nitrogen fixation under ambient conditions enabled by the heterojunctions of n-type Bi2MoO6 and oxygen-vacancy-rich p-type BiOBr. Nanoscale 11(21), 10439 (2019)

    Article  CAS  Google Scholar 

  52. J. Yang, Y. Guo, R. Jiang, F. Qin, H. Zhang, W. Lu, J. Wang, J.C. Yu, High-efficiency “working-in-tandem” nitrogen photofixation achieved by assembling plasmonic gold nanocrystals on ultrathin titania nanosheets. J. Am. Chem. Soc. 140(27), 8497 (2018)

    Article  CAS  Google Scholar 

  53. C. Xiao, H. Hu, X. Zhang, D.R. MacFarlane, Nanostructured gold/bismutite hybrid heterocatalysts for plasmon-enhanced photosynthesis of ammonia. ACS Sustain. Chem. Eng. 5(11), 10858 (2017)

    Article  CAS  Google Scholar 

  54. H. Li, J. Shang, J. Shi, K. Zhao, L. Zhang, Facet-dependent solar ammonia synthesis of BiOCl nanosheets via a proton-assisted electron transfer pathway. Nanoscale 8(4), 1986 (2016)

    Article  CAS  Google Scholar 

  55. Y. Bai, L. Ye, T. Chen, L. Wang, X. Shi, X. Zhang, D. Chen, Facet-dependent photocatalytic N2 fixation of bismuth-rich Bi5O7I nanosheets. ACS Appl. Mater. Interfaces 8(41), 27661 (2016)

    Article  CAS  Google Scholar 

  56. S. Wang, X. Hai, X. Ding, K. Chang, Y. Xiang, X. Meng, Z. Yang, H. Chen, J. Ye, Light-switchable oxygen vacancies in ultrafine Bi5O7Br nanotubes for boosting solar-driven nitrogen fixation in pure water. Adv. Mater. 29(31), 1701774 (2017)

    Article  CAS  Google Scholar 

  57. Y. Yang, T. Zhang, Z. Ge, Y. Lu, H. Chang, P. Xiao, R. Zhao, Y. Ma, Y. Chen, Highly enhanced stability and efficiency for atmospheric ammonia photocatalysis by hot electrons from a graphene composite catalyst with Al2O3. Carbon 124, 72 (2017)

    Article  CAS  Google Scholar 

  58. S. Sun, X. Li, W. Wang, L. Zhang, X. Sun, Photocatalytic robust solar energy reduction of dinitrogen to ammonia on ultrathin MoS2. Appl. Catal. B-Environ. 200, 323 (2017)

    Article  CAS  Google Scholar 

  59. Y. Zhao, Y. Zhao, G.I.N. Waterhouse et al., Layered-double-hydroxide nanosheets as efficient visible-light-driven photocatalysts for dinitrogen fixation. Adv. Mater. 29(42), 1703828 (2017)

    Article  CAS  Google Scholar 

  60. N. Zhang, A. Jalil, D. Wu et al., Refining defect states in W18O49 by Mo doping: A strategy for tuning N2 activation towards solar-driven nitrogen fixation. J. Am. Chem. Soc. 140(30), 9434 (2018)

    Article  CAS  Google Scholar 

  61. S. Sun, Q. An, W. Wang, L. Zhang, J. Liu, W.A.G. Iii, Efficient photocatalytic reduction of dinitrogen to ammonia on bismuth monoxide quantum dots. J. Mater. Chem. A 5(1), 201 (2017)

    Article  CAS  Google Scholar 

  62. X. Li, W. Wang, D. Jiang, S. Sun, L. Zhang, X. Sun, Efficient solar-driven nitrogen fixation over carbon–tungstic-acid hybrids. Chem. Eur. J. 22(39), 13819 (2016)

    Article  CAS  Google Scholar 

  63. S. Hu, X. Chen, Q. Li, F. Li, Z. Fan, H. Wang, Y. Wang, B. Zheng, G. Wu, Fe3+ doping promoted N2 photofixation ability of honeycombed graphitic carbon nitride: the experimental and density functional theory simulation analysis. Appl. Catal. B-Environ. 201, 58 (2017)

    Article  CAS  Google Scholar 

  64. L. Li, Y. Wang, S. Vanka, X. Mu, Z. Mi, C.J. Li, Nitrogen photofixation over III-nitride nanowires assisted by ruthenium clusters of low atomicity. Angew. Chem. Int. Ed. 56(30), 8701 (2017)

    Article  CAS  Google Scholar 

  65. S. Cao, N. Zhou, F. Gao, H. Chen, F. Jiang, All-solid-state Z-scheme 3,4-dihydroxybenzaldehyde-functionalized Ga2O3/graphitic carbon nitride photocatalyst with aromatic rings as electron mediators for visible-light photocatalytic nitrogen fixation. Appl. Catal. B-Environ. 218, 600 (2017)

    Article  CAS  Google Scholar 

  66. Q. Liu, L. Ai, J. Jiang, MXene-derived TiO2@C/g-C3N4 heterojunctions for highly efficient nitrogen photofixation. J. Mater. Chem. A 6(9), 4102 (2018)

    Article  CAS  Google Scholar 

  67. Y. Wang, W. Wei, M. Li, S. Hu, J. Zhang, R. Feng, In situ construction of Z-scheme g-C3N4/Mg1.1Al0.3Fe0.2O1.7 nanorod heterostructures with high N2 photofixation ability under visible light. RSC Adv.7(29), 18099 (2017)

    CAS  Google Scholar 

  68. C. Li, T. Wang, Z.J. Zhao, W. Yang, J.F. Li, A. Li, Z. Yang, G.A. Ozin, J. Gong, Promoted fixation of molecular nitrogen with surface oxygen vacancies on plasmon-enhanced TiO2 photoelectrodes. Angew. Chem. Int. Ed. 57(19), 5278 (2018)

    Article  CAS  Google Scholar 

  69. M. Lashgari, P. Zeinalkhani, Photocatalytic N2 conversion to ammonia using efficient nanostructured solar-energy-materials in aqueous media: a novel hydrogenation strategy and basic understanding of the phenomenon. Appl. Catal. A-Gen. 529, 91 (2017)

    Article  CAS  Google Scholar 

  70. Y. Hao, X. Dong, S. Zhai, H. Ma, X. Wang, X. Zhang, Hydrogenated bismuth molybdate nanoframe for efficient sunlight-driven nitrogen fixation from air. Chem. Eng. J. 22(52), 18722 (2016)

    CAS  Google Scholar 

  71. L. Zhou, C.E. Boyd, Comparison of Nessler, phenate, salicylate and ion selective electrode procedures for determination of total ammonia nitrogen in aquaculture. Aquaculture 450, 187 (2016)

    Article  CAS  Google Scholar 

  72. Y. Lin, H. Hu, Y.H. Hu, Role of ZnO morphology in its reduction and photocatalysis. Appl. Surf. Sci. 502, 144202 (2020)

    Article  CAS  Google Scholar 

  73. D.Y. Kim, T. Yoon, Y.J. Jang, J.H. Lee, Y. Na, B.J. Lee, J.S. Lee, K.S. Kim, Band gap narrowing of zinc orthogermanate by dimensional and defect modification. J. Phys. Chem. C 123(23), 14573 (2019)

    Article  CAS  Google Scholar 

  74. Y.J. Jang, I. Jeong, J. Lee, J. Lee, M.J. Ko, J.S. Lee, Unbiased sunlight-driven artificial photosynthesis of carbon monoxide from CO2 using a ZnTe-based photocathode and a perovskite solar cell in tandem. ACS Nano 10(7), 6980 (2016)

    Article  CAS  Google Scholar 

  75. Y.J. Jang, J. Ryu, D. Hong, S. Park, J.S. Lee, A multi-stacked hyperporous silicon flake for highly active solar hydrogen production. Chem. Commun. 52(67), 10221 (2016)

    Article  CAS  Google Scholar 

  76. J. Ryu, Y.J. Jang, S. Choi, H.J. Kang, H. Park, J.S. Lee, S. Park, All-in-one synthesis of mesoporous silicon nanosheets from natural clay and their applicability to hydrogen evolution. NPG Asia Mater. 8(e248), 1 (2016)

    CAS  Google Scholar 

  77. N.C. Kani, A. Prajapati, B.A. Collins, J.D. Goodpaster, M.R. Singh, Competing effects of pH, cation identity, H2O saturation, and N2 concentration on the activity and selectivity of electrochemical reduction of N2 to NH3 on electrodeposited Cu at ambient conditions. ACS Catal. 10(24), 14592 (2020)

    Article  CAS  Google Scholar 

  78. X. Wei, M. Pu, Y. Jin, M. Wessling, Efficient electrocatalytic N2 reduction on three-phase interface coupled in a three-compartment flow reactor for the ambient NH3 synthesis. ACS Appl. Mater. Interfaces 13(18), 21411 (2021)

    Article  CAS  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by the research fund of Hanyang University (HY202000000002708), the Korea Institute of Energy Technology Evaluation and Planning (KETEP) and the Ministry of Trade, Industry & Energy(MOTIE) (No. 20202020800330), the Graduate school of Post Plastic specialization of Korea Environmental Industry & Technology Institute grant funded by the Ministry of Environment of Republic of Korea and National Research Foundation (NRF) Grant funded by the Korea government (No. 2020R1G1A1100104, No. 2019M1A2A2065612, No. 2018R1A2A1A05077909).

Author information

Authors and Affiliations

Authors

Contributions

SC, SMK, YL, JS, JJ, and Prof. YJJ contributed to prepare this review. Prof. YJJ and Prof. JSL supervised this work and Prof. YJJ, Prof. JSL and SC are major contributors in writing the manuscript. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Jae Sung Lee or Youn Jeong Jang.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Choe, S., Kim, S.M., Lee, Y. et al. Rational design of photocatalysts for ammonia production from water and nitrogen gas. Nano Convergence 8, 22 (2021). https://doi.org/10.1186/s40580-021-00273-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40580-021-00273-8

Keywords