Two-dimensional building blocks for photocatalytic ammonia production

Jingrun Ran *, Bingquan Xia , Yanzhao Zhang and Shi-Zhang Qiao *
School of Chemical Engineering and Advanced Materials, The University of Adelaide, Adelaide, SA 5005, Australia. E-mail: s.qiao@adelaide.edu.au; Jingrun.ran@adelaide.edu.au

Received 17th November 2020 , Accepted 6th January 2021

First published on 6th January 2021


Abstract

Owing to its extensive utilization in fertilizer generation and as an energy carrier, ammonia (NH3) is deemed to be one of the most essential chemicals. Currently, NH3 generation mostly depends on the Haber–Bosch approach under harsh conditions, resulting in tremendous energy usage and environmental problems. Photocatalytic NH3 generation represents a clean, inexpensive and environmentally friendly method to transform water and nitrogen into ammonia utilizing sunlight under ambient conditions. Recently, two-dimensional (2D) building blocks have received great attention in the photocatalysis field thanks to their outstanding features of high surface area, plentiful reactive sites, ultrathin thickness and short charge-to-surface transfer distance. This perspective summarizes the design and synthesis of photocatalysts prepared utilizing 2D building blocks towards light-driven NH3 production. Our contribution highlights the in-depth and comprehensive structure/composition–performance relationship in 2D building block based photocatalysts for light-induced NH3 production. We also discuss the delicate and insightful reaction mechanisms in 2D building block based photocatalytic NH3 production. Finally, we propose the possible opportunities in merging advanced characterization techniques as well as powerful theoretical computations towards the rational design and fabrication of high-performance 2D material based photocatalysts towards light-induced NH3 generation.


image file: d0ta11201a-p1.tif

Jingrun Ran

Dr Jingrun Ran received his BE and ME degrees in Materials Science and Engineering from Wuhan University of Technology, and PhD degree in Chemical Engineering from the University of Adelaide. Now he is working as an ARC DECRA Fellow in Prof. Shi-Zhang Qiao's group, focusing on the atomic-level design and synthesis of photocatalysts for producing energy fuels and value-added chemicals using renewable solar energy. Dr Jingrun Ran has been recognized as a Clarivate Highly Cited Researcher in 2020.

image file: d0ta11201a-p2.tif

Bingquan Xia

Mr Bingquan Xia received his Master's degree in chemistry after graduating from Wuhan University in 2016. He is currently a PhD candidate under the supervision of Prof. Shi-Zhang Qiao at the University of Adelaide. His research is focused on the development of highly efficient photocatalysts for solar energy conversion.

image file: d0ta11201a-p3.tif

Yanzhao Zhang

Yanzhao Zhang received his BE degree from Wuhan University and ME degree in Materials Science and Engineering from Zhejiang University, and is now a PhD candidate under the supervision of Prof. Shi-Zhang Qiao and Dr Jingrun Ran at the University of Adelaide. Presently, Yanzhao is working on photocatalytic CO2 conversion.

image file: d0ta11201a-p4.tif

Shi-Zhang Qiao

Shi-Zhang Qiao received his PhD degree in chemical engineering from the Hong Kong University of Science and Technology in 2000, and is currently a Chair Professor at the School of Chemical Engineering and Advanced Materials at the University of Adelaide, Australia. His research expertise is in nanomaterials for new energy technologies. He has co-authored more than 430 papers in refereed journals with over 73[thin space (1/6-em)]000 citations and an h-index of 141. In recognition of his achievements in research, he was honoured with the prestigious ARC Laureate Fellow (2017), ExxonMobil Award (2016), ARC Discovery Outstanding Researcher Award (2013), and the Emerging Researcher Award (2013, the ENFL Division of the American Chemical Society).


1. Introduction

As a pivotal feedstock for synthesizing fertilizers and an important energy carrier, ammonia (NH3) is produced in an amount of ca. 150 million tons annually worldwide.1–23 Presently, industrial-scale preparation of NH3 still depends upon the conventional Haber–Bosch approach developed in the early 1900s. But this Haber–Bosch process is performed under harsh conditions: elevated temperature (350–550 °C) together with high pressure (150–350 atmosphere pressure) using iron-based catalysts, which utilizes ca. ∼1% of global energy supply and large amounts of fossil fuels to produce hydrogen (H2) as a feedstock.1–23 Hence, the development of a sustainable and environmentally benign strategy to produce NH3 is of paramount significance.

Photocatalytic NH3 production is a clean, cost-effective, and eco-friendly technique capable of converting nitrogen (N2) and water (H2O) into NH3 using renewable solar energy at room temperature and under atmospheric pressure.24–51 To realize industrial-scale photocatalytic NH3 production, the core target is the exploration of highly active, strongly reliable and inexpensive nanostructured materials. Recently, two-dimensional (2D) materials have garnered tremendous attention in a variety of fields (e.g., electronics, catalysis and optoelectronics) ever since the discovery of graphene.24–63 Their distinctive features of high surface area, abundant reactive centres, ultrathin thickness and short bulk-to-surface distance make them outstanding building blocks to construct high-performance photocatalysts.

A range of 2D building blocks, e.g., graphitic carbon nitride (g-C3N4),24–27 BiOCl,30 Bi3O4Br,31 BiOBr,32 layered double hydroxides (LDHs),33–35 TiO2,36 MoS2,38 and SmOCl,39 have been developed as single-component 2D photocatalysts for NH3 production. Accordingly, a variety of strategies, e.g., elemental doping,27,30,33,35,36 creating vacancies,24,25,27,30–37,39 modification on the termination,24,26 producing porosity,24,27 and/or crystal facet engineering,30,32 have been adopted to tailor the physicochemical characteristics of the above 2D building blocks for achieving enhanced activity, selectivity and stability in photocatalytic NH3 generation. Besides, heterojunctions based on 2D building blocks have also been designed and prepared for light-driven NH3 production. A series of binary heterojunctions, i.e., zero-dimensional (0D)/2D,40–46 2D/2D47,48 and three-dimensional (3D)/2D,49 as well as ternary heterostructures50,51 have been explored as efficient, highly selective and robust photocatalysts towards light-induced generation of NH3.

In this perspective, we for the first time summarize all the photocatalysts synthesized utilizing 2D building blocks towards photocatalytic NH3 production. The in-depth and overall structure/composition–performance relationship in these 2D building block based photocatalysts is discussed. Additionally, the insightful and delicate reaction mechanisms in photocatalytic NH3 production are also explained. Finally, we propose the possible opportunities in this research field with special focus on merging the advanced characterization techniques, e.g., aberration-corrected scanning transmission electron microscopy (AC-STEM), synchrotron-based X-ray absorption spectroscopy (XAS), in situ Raman, and in situ Fourier transform infrared (FTIR) spectroscopy, and powerful theoretical calculations, to develop high-performance photocatalysts based on 2D building blocks for NH3 generation.

2. Merits of 2D building blocks in photocatalytic NH3 production

The distinct physicochemical features of 2D building blocks endow them with many outstanding merits in photocatalytic NH3 production:24–51 (i) their large surface area and abundant reactive sites facilitate the adsorption/activation/reduction of N2 into NH3; (ii) their ultrathin thickness benefits the dissociation and migration of photo-induced charge carriers from the bulk onto the surface; (iii) their large surface area facilitates the formation of electronic coupling with other materials for efficient interfacial charge carrier separation and migration; (iv) their tailorable thickness allows the alteration of band gap width via the quantum confinement effect, accompanied by modulation of light absorption capacity and conduction/valence band edge positions; (v) their highly exposed surface atoms also facilitate the adoption of various engineering strategies (e.g., doping, creating vacancies and single-atom anchoring) to acquire the desired properties and functions.

It should be noted that various types of active sites, e.g., cation/anion vacancies,25,27,30,31,33–37,39,42,45 incorporated heteroatoms,35,49 and anchored single atoms,32,40,41 have been developed on 2D material based photocatalysts for efficient adsorption, activation and reduction of N2 to produce NH3. Each of the above types alone or two types of active sites together could apparently promote the adsorption/activation/reduction of N2 molecules, as confirmed by experimental characterization, e.g., N2 temperature-programmed desorption (TPD) and/or theoretical computations, thus leading to apparently improved activity and selectivity in photocatalytic N2-to-NH3 conversion.

3. Single-component 2D photocatalysts

Up to now, various single-component 2D photocatalysts have been developed and utilized in photocatalytic NH3 generation (see Table 1).24–27,30–39 A series of approaches, e.g., heteroatom incorporation, generation of vacancies, morphology control and surface modifications, have been explored to engineer the properties of these single-component 2D photocatalysts for highly efficient, highly selective and steady light-driven NH3 generation. They are classified into four categories: (1) g-C3N4 based 2D photocatalysts;24–27 (2) bismuth based 2D photocatalysts;30–32 (3) LDH based 2D photocatalysts;33–35 and (4) other 2D photocatalysts,36–39 which are introduced in detail as follows.
Table 1 Single-component 2D photocatalysts for NH3 production
Photocatalyst Synthesis method Light source Reactant solution NH3 production References (year)
Activity (μmol h−1 g−1) Quantum efficiency (%) Stability Detection method
Porous g-C3N4 with nitrogen vacancies and cyano groups Alkali-assisted heat treatment Simulated solar light (AM 1.5G, 100 mW cm−2) 20 v% methanol aqueous solution 1590 Nessler's reagent 24 (2019)
g-C3N4 with carbon vacancies High-temperature peeling method Xe lamp Water 84 >100 min Nessler's reagent 25 (2019)
g-C3N4 with cyano groups and intercalated K+ KOH-assisted calcination and ultrasonication Xe lamp (λ > 400 nm) Ethylene glycol 3420 >4 hours Nessler's reagent 26 (2019)
S doped g-C3N4 with C vacancies Thermal polymerization of thiourea Xe lamp 4 v% methanol aqueous solution 5990 >4 hours Nessler's reagent 27 (2018)
BiOCl with Br doping and O vacancies Surfactant assisted solvothermal approach Xe lamp (λ > 400 nm) Deionized water 126 >5 hours Nessler's reagent 30 (2019)
Bi3O4Br with Bi and O vacancies Surfactant assisted self-assembly strategy Xe lamp Distilled water 50.8 >4 hours Nessler's reagent 31 (2019)
BiOBr with O vacancies Surfactant assisted hydrothermal approach Xe lamp Deionized water 54.7 >12 hours Nessler's reagent 32 (2018)
CuCr LDH with O vacancies and doped Cu2+ ions Co-precipitation method Xe lamp (λ > 400 nm) Double distilled water 57.1 0.10 at 500 nm >5 hours Nessler's reagent 33 (2017)
ZnCr LDH with oxygen and cation vacancies Alkali-etching and hydrothermal approach Xe lamp Ultra-pure water 33.19 0.95 at 380 nm, 0.34 at 420 nm, 0.11 at 550 nm >5 hours Indophenol blue method 34 (2020)
ZnAl LDH with O vacancies and doped Cuδ+ Co-precipitation approach Xe lamp Ultra-pure water 110 1.77 at 265 nm, 0.56 at 365 nm >40 hours Indophenol blue method 35 (2020)
TiO2 with O vacancies and doped Cu Hydrothermal approach Xe lamp Ultra-pure water 78.9 0.08 at 600 nm, 0.05 at 700 nm >5 hours Indophenol blue method 36 (2019)
MoO3−x with O vacancies Hydrothermal approach Xe lamp Distilled water 1.11 0.013 at 365 nm >24 hours Ion chromatography 37 (2019)
MoS2 Hydrothermal approach Xe lamp (λ > 420 nm) Deionized water 325 >10 hours Indophenol blue method 38 (2017)
SmOCl with O vacancies Wet-chemical method using graphene oxide as the template Xe lamp Deionized water 426 0.32 at 420 nm >4 hours Nessler's reagent 39 (2019)


3.1 g-C3N4 based 2D photocatalysts

In recent years, g-C3N4 has attracted enormous attention in the photocatalysis field thanks to its appealing properties of appropriate band gap width, favourable band edge positions, reliable stability, earth abundance and excellent processability.24–27,62,63 Particularly, several g-C3N4 based 2D photocatalysts are designed and prepared towards photocatalytic NH3 generation.24–27 Various techniques, including elemental doping,27 vacancy defect creation,24,25,27 terminal modification24,26 and/or porosity generation,24,27 have been applied to engineer the physicochemical characteristics and optimize the photocatalytic NH3 production performance of 2D g-C3N4.

Xue et al.24 reported the synthesis of porous g-C3N4 nanosheets with cyano groups and nitrogen defects by alkali-assisted calcination of urea. The as-prepared porous g-C3N4 nanosheets with cyano groups and nitrogen defects present an enhanced photocatalytic N2 fixation rate of 1590 μmol h−1 g−1 in comparison to pure g-C3N4 without any co-catalyst under simulated solar irradiation (AM 1.5G). The elevated photocatalytic activity is caused by the synergistic impacts of strengthened light harvesting, advanced photo-induced charge carrier migration and adsorption of N2 gas molecules. In another study, Zhang et al.25 synthesized ultrathin g-C3N4 nanosheets with abundant carbon vacancies on the surface via the thermal exfoliation route. The as-prepared ultrathin g-C3N4 nanosheets display a photocatalytic NH4+ production rate of 54 μmol L−1 in 100 min in the presence of no co-catalysts or sacrificial reagents, about 2.25 times larger than that of bulk g-C3N4. Such an improved activity originates from the carbon vacancies on the surface and ultra-small thickness, which favour the efficient charge carrier dissociation and transfer both in the bulk and on the surface. Wang et al.26 prepared g-C3N4 nanoribbons with intercalated K+ and cyano groups (mCNN). Compared to the scanning electron microscopy (SEM) (see Fig. 1a) and transmission electron microscopy (TEM) (see Fig. 1a inset) images of pristine g-C3N4, the SEM (Fig. 1b) and TEM (Fig. 1b inset) images of mCNN display its ribbon-shaped structures with uniform size, due to the scissoring out effect of molten KOH in the preparation process. A visible-light-driven photocatalytic NH3 production activity of 3.42 mmol g−1 h−1 was obtained on the as-prepared mCNN, obviously larger than that on pristine g-C3N4 (1.11 mmol g−1 h−1) as displayed in Fig. 1c. Fig. 1d presents the photocatalytic NH3 production amounts of pristine g-C3N4 and mCNN photocatalysts utilizing N2 or Ar as the feed gas. As Ar was utilized as the feed gas without N2, pristine g-C3N4 and mCNN photocatalysts present average photocatalytic NH3 production activities of 0.23 and 0.52 mmol g−1 h−1, respectively. These results indicate that the nitrogen source of the produced NH3 arises from photocatalysts, as Ar is utilized as the feed gas. Both experiments and theoretical computations support that the cyano group in mCNN is regenerated with the assistance of intercalated K+, similar to the Mars–van Krevelen process. The regeneration of the cyano group results in the improved activity and stability of mCNN. In another study, g-C3N4 porous nanosheets with doped sulphur and carbon vacancies (SCNNSs) have been prepared via heating melamine and trithiocyanuric acid together.27 A photocatalytic NH3 generation activity of 5.99 mM h−1 g−1 was acquired on the as-synthesized SCNNSs using simulated sunlight illumination. This activity is 280% times larger than that of bulk S doped g-C3N4. It was found that the porous sheet structure, ultra-small thickness, and incorporation of sulphur and carbon vacancies induce better photocatalytic performance.


image file: d0ta11201a-f1.tif
Fig. 1 (a) SEM and TEM (inset) results of pristine g-C3N4. (b) SEM and TEM (inset) results of mCNN. (c) Photocatalytic NH3 production amounts of g-C3N4 and mCNN. (d) Photocatalytic NH3 production amounts versus illumination time plots of g-C3N4 and mCNN using N2 and Ar as the feed gases. Reproduced from ref. 26 with permission from John Wiley and Sons. (e) TEM image of Bi3O4Br NSs with abundant defects. (f) Atomic-resolution HAADF-STEM images of Bi3O4Br NSs with abundant defects; the image has been processed utilizing a Gaussian filter to eliminate noise. (g) Structural model displaying Bi defects on the surface in (f). (h) Line profiles of Bi3O4Br NSs with abundant defects. (i) Positron annihilation lifetime spectra of Bi3O4Br NSs with abundant defects and Bi3O4Br with few defects. (j) Photocatalytic NH3 production under light illumination. (k) Stability test of photocatalytic NH3 production of defect-rich Bi3O4Br NSs under light illumination. Reproduced from ref. 31 with permission from John Wiley and Sons.

Apart from the above studies, investigations based on theoretical calculations have been applied to study the g-C3N4 photocatalyst for NH3 production. For instance, Lv et al.28 utilized density functional theory (DFT) based computations to explore the single B atom loaded on holey graphitic carbon nitride (B@g-CN) towards the photocatalytic nitrogen reduction reaction (NRR). The DFT computation results indicate that efficient N2-to-NH3 conversion can be achieved via the enzymatic pathway on B@g-CN with a very low activation barrier of 0.61 eV and overpotential of 0.15 V. These values are lower than those of many metal-based catalysts. Besides, the photo-excited electrons on B@g-CN can sufficiently enable the NRR against the hydrogen evolution reaction (HER). In another study, on the basis of DFT calculation results, Ren et al.29 found that nitrogen vacancies not only boost the dissociation of photo-excited electron–hole pairs in g-C3N4, but also increase the light absorption. Besides, they also found that the corrugated configuration structure of g-C3N4 favours the N2 adsorption ability, thus improving the photocatalytic N2 fixation activity.

3.2 Bismuth based 2D photocatalysts

Due to the fascinating features of adjustable light-harvesting ability, internal electric field, easy creation of surface oxygen vacancy and photochemical stability, 2D Bi-based photocatalysts have drawn tremendous attention recently.30–32 Different 2D Bi-based photocatalysts, e.g., BiOCl,30 Bi3O4Br31 and BiOBr,32 were prepared and adopted for photocatalytic NH3 production. For example, Wu et al.30 prepared Br doped BiOCl micro-sheets with abundant O vacancies and exposed {001} facets (Br–BiOCl–OV) using a surfactant assisted solvothermal approach and a subsequent ion-exchange method. An NH3 production activity of 6.3 μmol h−1 was achieved on the optimized Br–BiOCl–OV utilizing visible light, ca. 150% larger than that of bare BiOCl. The incorporation of the Br ion not only leads to a better N2 activation effect based on theoretical calculations, but also narrows the band gap width as well as raising the conduction band position, resulting in enhanced light absorption and stronger reduction capacity. Besides, Br–BiOCl–OV also presents improved dissociation of photo-induced excitons. The aforementioned effects synergistically cause the elevated photocatalytic NH3 generation rate of Br–BiOCl–OV.

Di et al.31 synthesized single-unit-cell (SUC) Bi3O4Br nanosheets with confined defects using a polyvinylpyrrolidone (PVP) self-assembly approach. The TEM image of SUC Bi3O4Br nanosheets is displayed in Fig. 1e, confirming their 2D ultrathin nanosheet structure. Abundant point defects are observed on SUC Bi3O4Br nanosheets as shown in the aberration-corrected atomic-resolution high angle annular dark field scanning transmission electron microscopy (HAADF-STEM) image (Fig. 1f). The structure model in Fig. 1g presents the surface point defects in Fig. 1f. Accordingly, the line profiles in Fig. 1h indicate that the numerous point vacancies correspond to Bi vacancies confined in SUC Bi3O4Br NSs. The presence of Bi vacancy can significantly increase the production of oxygen vacancies as confirmed by both theoretical computations and experimental results. Fig. 1i exhibits the positron annihilation lifetime spectra of SUC Bi3O4Br NSs with abundant defects and Bi3O4Br with few defects. The acquired positron annihilation lifetime parameters indicate the higher concentration of Bi surface defects in defect-rich SUC Bi3O4Br NSs compared with that of defect-deficient Bi3O4Br NSs. As displayed in Fig. 1j, after one-hour reaction, the defect-rich SUC Bi3O4Br NSs exhibit a larger amount of NH3 (25.4 μmol L−1) in a N2 atmosphere, about 9.2 and 30.9 times larger than that of Bi3O4Br NSs with deficient defects and bulk Bi3O4Br, respectively. Besides, no apparent decrease of the photocatalytic NH3 production activity is observed after four cycles (Fig. 1k), suggesting the good stability of defect-rich SUC Bi3O4Br NSs. The enhanced photocatalytic NH3 production performance arises from the efficient dissociation and transportation of photo-excited electrons and holes, owing to their SUC configuration and surface defects. Moreover, Xue et al.32 fabricated oxygen vacancy engineered BiOBr ultrathin NSs mainly exposed with {001} crystal facets through utilizing PVP in the hydrothermal procedure. The BiOBr ultrathin NSs with oxygen vacancies exhibit about 10 times larger photocatalytic NH3 production rate (54.70 μmol g−1 h−1) compared to BiOBr nanoplates in the absence of any oxygen vacancies (5.75 μmol g−1 h−1). They found that the engineering of oxygen vacancies in BiOBr ultrathin NSs not only reduces the band gap and strengthens the light harvesting, but also raises the conduction band position for stronger reduction capability of photo-induced electrons. Besides, both the experiments and theoretical computations indicate that the adsorption and activation of N2 molecules could be promoted by oxygen vacancies.

3.3 LDH based 2D photocatalysts

As a novel category of metal-to-metal charge-transfer (MMCT) based system, LDHs are highly promising in the photocatalysis field, owing to their adjustable metal cation components and tailorable thickness combined with engineering of defects and band structures.33 Thus, a range of LDH based photocatalysts have been developed.33–35 For instance, Zhao et al.34 reported the fabrication of LDH NSs (i.e., ZnCr–LDH, ZnAl–LDH and NiAl–LDH) with plentiful O vacancies, cation vacancies and coordinatively unsaturated metal sites via an alkali-etching process. All the alkali-etched LDH NSs exhibit superior photocatalytic NH3 production activities compared to their un-etched counterparts. In particular, etched ZnCr–LDH presents a photocatalytic NH3 production activity of 33.19 μmol g−1 h−1 with an apparent quantum efficiency (AQE) of 0.11% at 550 nm, ca. 1000% times larger in contrast to that of un-etched ZnCr–LDH (3.15 μmol g−1 h−1). The alkali-etching approach not only elevates the conduction band edge of ZnCr–LDH to improve the reduction capacity of photo-induced electrons, but also reduces its band gap for stronger absorption of visible light. Besides, the alkali-etching approach generated vacancies in ZnCr–LDH, which act as trapping centres for photo-induced electrons and elongate the charge-carrier lifetimes. Moreover, the unsaturated Zn sites created by the alkali-etching method further favour the adsorption and activation of N2. Hence, boosted photocatalytic NH3 production activities are achieved on the etched ZnCr–LDH. In another report, a facile co-precipitation method was employed to synthesize a range of MIIMIII (MII = Mg, Zn, Ni, Cu; MIII = Al, Cr) LDH ultrathin NSs.33Fig. 2a and b show the TEM images of CuCr LDH ultrathin NSs, suggesting their NS structure with an averaged lateral size of about 20 nm and a thickness of approximately 2.5 nm. The high-resolution (HR)-TEM image of CuCr LDH ultrathin NSs (Fig. 2c) displays a d spacing of 0.24 nm, ascribed to the (009) facet of CuCr LDH ultrathin NSs. As presented in Fig. 2d, the averaged length of the first Cr–O shell within CuCr-NS is 1.989 Å, smaller than that acquired within CuCr-bulk. Furthermore, the Cr in CuCr-NS shows a decreased coordination number of 5.5, in comparison to the Cr in CuCr-bulk (6.0). This result suggests a seriously distorted structure around Cr cations within CuCr-NS, in agreement with the presence of plentiful oxygen vacancies. Fig. 2e shows the explanation of the in-plane compressive strain caused by the abundant oxygen vacancies. As presented in Fig. 2f, many LDH NSs exhibit photocatalytic NH3 production activities utilizing visible light (λ > 400 nm). Particularly, CuCr–LDH NSs display the highest photocatalytic NH3 generation rate of 142.9 μmol L−1. Additionally, an apparent quantum yield (AQY) of 0.10% at 500 nm was also obtained on CuCr–LDH NSs. To study the adsorption/activation/reduction of N2, in situ diffuse reflectance infrared Fourier transform spectroscopy was adopted to probe the reaction intermediates on the CuCr–LDH NS surface (Fig. 2g). The characteristic bands at 1661, 1557 and 1448 cm−1 correspond to the antisymmetric as well as symmetric deformations of surface NH4+ species. Both the adsorption/activation of N2 and H2O molecules are promoted via the abundant O vacancies in CuCr–LDH NSs. In addition, the structure distortions and compressive strain caused by the incorporation of Cu2+ also result in enhanced interaction between N2 and LDH, thus increasing NH3 production. Furthermore, Zhang et al.35 have prepared ZnAl–LDH NS with incorporated Cu2+ and plentiful oxygen vacancies utilizing an easy co-precipitation route. The Cu modified ZnAl–LDH NS exhibits an outstanding photocatalytic NH3 production activity of 110 μmol g−1 h−1 in pure water using UV-visible illumination. Excellent robustness in photocatalytic NH3 production was also found on this Cu modified ZnAl–LDH NS. Its outstanding photocatalytic performance arises from the presence of numerous oxygen vacancies and electron-rich Cuδ+ advancing the N2 adsorption/activation as well as electron–hole dissociation and transportation.
image file: d0ta11201a-f2.tif
Fig. 2 (a and b) TEM and (c) HRTEM results of CuCr-NS. (d) Magnitude of k2-weighted FT of Cr K-edge EXAFS spectra of CuCr-bulk and CuCr-NS. (e) Explanation of the in-plane biaxial compressive strain in the as-prepared LDH nanosheets. (f) Photocatalytic NH3 production amount of various LDH photocatalysts using visible-light irradiation (λ > 400 nm). (g) In situ IR spectra collected on CuCr-NS in the process of 125 min UV-vis irradiation in a N2 and water vapor atmosphere. Reproduced from ref. 33 with permission from John Wiley and Sons. (h) Magnitude of k2-weighted FT of Ti K-edge EXAFS spectra for X%-TiO2 NS (X = 0, 1, 3, 6, 8) and bulk-TiO2 NSs; the inset in (h) displays the amplified view of the Ti–O signal. (i) 2D structure model for TiO2 NSs with oxygen vacancies and engineered strain. (j) Photocatalytic NH3 production amounts for various samples using water as the proton source after 1 hour UV-vis illumination; the produced NH3 was detected using Nessler's reagent. (k) Time course of photocatalytic NH3 production on 6%-TiO2 utilizing 600 nm and 700 nm illumination; the produced NH3 was detected by ion chromatography. Reproduced from ref. 36 with permission from John Wiley and Sons.

3.4 Other 2D photocatalysts

Apart from the above-mentioned 2D photocatalysts, other 2D photocatalysts, such as TiO2 NSs,36 MoO3−x nanobelts,37 ultrathin MoS2 NSs38 and amorphous SmOCl NSs,39 have also been explored for photocatalytic NH3 generation. For example, Zhao et al.36 doped copper in TiO2 NSs for producing more oxygen vacancies and massive compressive strain. X-ray absorption fine structure (XAFS) was adopted to study the local atomic structure and cation coordination in X%-TiO2 (X = 0, 1, 3, 6, 8) resulting from Cu incorporation in the TiO2 NSs. Fig. 2h presents the magnitude of k2-weighted Fourier transforms of Ti K-edge EXAFS for X%-TiO2 (X = 0, 1, 3, 6, 8) and bulk-TiO2. It is observed that the Ti–O shell peak intensity is increasingly lowered upon the increase of Cu concentration in TiO2 from 0% to 6%, in agreement with the increase of O vacancies. Fig. 2i demonstrates the strain caused by the O vacancies and Jahn–Teller effect arising from the incorporation of Cu. As can be seen from Fig. 2j, 6%-TiO2 presents the largest photocatalytic NH3 generation activity of 78.9 μmol g−1 h−1 among various samples. Besides, Fig. 2k shows that no NH3 was produced in an Ar atmosphere using 600 nm and 700 nm illumination. In contrast, in a N2 flow, the concentration of generated NH3 rises linearly with the illumination time. 6%-TiO2 exhibits photocatalytic NH3 production activities of 1.54 and 0.72 μmol g−1 h−1 utilizing 600 nm and 700 nm illumination, respectively. Moreover, 6%-TiO2 shows the excellent reliability of photocatalytic NH3 production in 5 cycles of testing. The outstanding photocatalytic NH3 production performance is ascribed to the ample oxygen vacancies and compressive strain arising from the Jahn–Teller distortion via doping Cu, which are supported by a series of characterization techniques, e.g., X-ray diffraction (XRD), XAFS and electron paramagnetic resonance (EPR) spectroscopy, together with theoretical calculations. Moreover, the preparation of oxygen vacancy-rich MoO3−x nanobelts was accomplished through adopting a hydrothermal approach.37 The oxygen vacancies are found on the (001) and (100) crystal facets of MoO3−x nanobelts confirmed by the scanning transmission electron microscopy (STEM) characterizations results. Efficient photocatalytic generation of NH3via N2 reduction was realized on oxygen vacancy-rich MoO3−x nanobelts, thanks to the existence of oxygen vacancies benefiting the chemisorption/activation of N2 molecules.

The preparation of ultrathin MoS2 using a hydrothermal method was presented by Sun et al.38 The ultrathin MoS2 exhibits a photocatalytic NH3 generation activity of 325 μmol h−1 g−1 utilizing no electron donor or co-catalyst. This excellent photocatalytic activity of ultrathin MoS2 arises from the light-induced trions activating and converting N2 molecules into NH3 through a simultaneous six-electron reduction procedure.

Hou et al.39 have prepared amorphous SmOCl nanosheets (A-SmOCl) utilizing graphene oxide (GO) as a template by a wet-chemical approach. The as-fabricated A-SmOCl displays a photocatalytic NH3 generation rate of 426 μmol h−1 g−1 under xenon light illumination (320–780 nm), with an AQY efficiency of 0.32% at 420 nm. Its impressive photocatalytic activity is ascribed to the presence of plentiful O vacancies in A-SmOCl enhancing the N2 adsorption/activation and increasing the light absorption range. Besides, the O K-edge XAS result corroborates the strengthened Sm–O covalency, which advances the migration of photo-induced electrons to the chemisorbed N2 molecules, thus increasing the photocatalytic activity of A-SmOCl.

For most single-component 2D photocatalysts discussed in Sections 3.1–3.4, the generation of vacancies (e.g., O vacancies, N vacancies and C vacancies)24,25,27,30–37,39 is a general and effective strategy to apparently boost the adsorption of N2 molecules and promote their subsequent activation and reduction. Thus, the light-induced N2-to-NH3 conversion performance is greatly enhanced on these single-component 2D photocatalysts engineered with vacancies.

4. 2D material based heterostructured photocatalysts

The unique properties (e.g., large surface area and highly exposed surface atoms) of 2D building blocks benefit their electronic coupling with other materials to achieve efficient interfacial dissociation and transportation of photo-induced charge carriers. In addition, a robust combination can also be achieved to realize high stability in photocatalytic NH3 production. Hence, various 2D material based heterostructured photocatalysts have been designed and synthesized towards photocatalytic N2 reduction (see Table 2).40–51 They can be categorized into 0D/2D binary heterostructures,40–46 2D/2D binary heterostructures,47,48 3D/2D binary heterostructures49 and ternary heterostructures.50,51
Table 2 Heterostructured 2D photocatalysts for NH3 production
Photocatalyst Synthesis method Light source Reactant solution NH3 production References (year)
Activity (μmol h−1 g−1) Quantum efficiency (%) Stability Detection method
0D/2D Ru/TiO2 Impregnation and calcination in H2/Ar Xe lamp 20 vol% ethanol aqueous solution ca. 3.3 >2 hours Indophenol blue method 40 (2019)
0D/2D Cu/g-C3N4 Impregnation and annealing treatment in N2 Xe lamp (780 nm > λ > 420 nm) 20 vol% ethanol aqueous solution 186 1.01 at 420 nm >3 hours Nessler's reagent 41 (2018)
0D/2D Au/TiO2 Self-assembly via electrostatic attraction Xe lamp (λ > 420 nm) 10 vol% methanol aqueous solution 130.5 0.82 at 550 nm >9 hours Indophenol blue method 42 (2018)
0D/2D Au/(BiO)2CO3 Hydrothermal method and chemical bath deposition Xe lamp Milli-Q water 38.23 <1 hour Indophenol blue method 43 (2017)
0D/2D Agln2S4/MXene Ti2C3 Hydrothermal method Xe lamp (λ > 400 nm) 20 v% methanol aqueous solution 38.8 0.07 at 420 nm <15 hours Indophenol blue method 44 (2019)
0D/2D AgCl/δ-Bi2O3 Hydrothermal controllable precipitation approach Xe lamp (λ > 400 nm) Deionized water 606 >3 hours Nessler's reagent 45 (2019)
0D/2D POM/RGO Stirring at 85 °C using L-ascorbic as the reducing reagent Xe lamp Distilled water 130.3 μmol L−1 h−1 >1 hour Nessler's reagent 46 (2019)
2D/2D MoO2/BiOCl Mechanical mixing Xe lamp Deionized water 35 >5 hours Nessler's reagent 47 (2019)
2D/2D BP/g-C3N4 Ultrasonication, drying and calcination Xe lamp (λ > 420 nm) 5 vol% methanol aqueous solution 347.5 μmol L−1 h−1 >20 hours Nessler's reagent 48 (2018)
3D/2D PrCO3OH/g-C3N4 Hydrothermal method Xe lamp 10 vol% methanol aqueous solution 8900 >5 hours Nessler's reagent 49 (2018)
Ternary MoS2/C–ZnO Hydrothermal, calcination and photo-deposition Xe lamp (λ > 420 nm) 5 vol% ethanol aqueous solution 245.7 μmol L−1 g−1 h−1 >24 hours Nessler's reagent 50 (2018)
Ternary TiO2@C/g-C3N4 One-step calcination Xe lamp (λ > 420 nm) 20 vol% methanol aqueous solution 250.6 0.14 at 420 nm >10 hours Nessler's reagent 51 (2019)


4.1 0D/2D binary heterostructures

0D/2D heterostructured photocatalysts are the most extensively studied binary heterostructured photocatalysts. Based on the intrinsic features of the formed heterojunction, they are classified into metal/semiconductor 0D/2D heterostructures,40–43 semiconductor/semiconductor 0D/2D heterostructures,44,45 and semiconductor/semimetal 0D/2D heterostructures.46

For instance, Liu et al.40 anchored single-atom Ru on TiO2 NS by impregnation and calcination in an Ar–H2 atmosphere. Both the aberration-corrected HAADF-STEM and synchrotron-based EXAFS spectra confirm the existence of single-atom Ru on the TiO2 NS surface. The DFT calculations further reveal the largest adsorption energy of the Ru–O–Ru structure in defective TiO2 with O vacancies, indicating that the formation of O vacancies via thermal H2 treatment could stabilize the atomic dispersion of Ru on the TiO2 NS surface. The single-atom Ru anchored TiO2 NSs exhibit a higher photocatalytic NH3 production activity (56.3 μg h−1 g−1) than unloaded TiO2 NSs (22.2 μg h−1 g−1) utilizing xenon light. This elevated photocatalytic NH3 production rate arises from the presence of single-atom Ru facilitating the charge separation/migration via accepting the photo-induced electrons into the empty d orbitals as well as enhancing the N2 adsorption and activation. In another study, single-atom Cu anchored g-C3N4 was fabricated by Huang et al.41 The as-prepared Cu loaded g-C3N4 displays a photocatalytic NH3 generation rate of 186 μmol h−1 g−1 utilizing visible-light illumination (λ > 420 nm) and a quantum efficiency of 1.01% at 420 nm. Such an activity is ca. 8 times larger than that of unmodified g-C3N4. The improved photocatalytic performance is ascribed to the active isolated π electrons as well as excellent adsorption capacity on the positively charged Cu ions, as corroborated by EXAFS, operando FTIR, operando EPR and theoretical computation results. Moreover, Yang et al.42 reported the preparation of Au nanosphere loaded TiO2 ultrathin nanosheets (UNSs) with oxygen vacancies (Au/TiO2–OV). Fig. 3a shows that Au nanospheres are uniformly dispersed on the surface of TiO2 UNSs with O vacancies (TiO2–OV). The existence of OVs was tested by low-temperature EPR analysis. TiO2 and Au/TiO2 exhibit no EPR signals (Fig. 3b). TiO2–OV and Au/TiO2–OV display a peculiar OV signal with a g factor of 1.998, indicating the existence of OVs (Fig. 3b). N2 TPD was applied to analyze the N2 adsorption ability of all the as-prepared samples. Both TiO2 and Au/TiO2 samples merely display an adsorption peak, resulting from N2 physisorption. In comparison, apart from the physisorption peak, both TiO2–OV and Au/TiO2–OV exhibit a peak located at a higher temperature, attributed to the N2 chemisorption (Fig. 3c). This result indicates that N2 chemisorption occurs at the OV sites on the TiO2 NS surface. The electron migration from OV-induced Ti3+ and N2 is deemed to be the major cause of the N2 chemisorption. As shown in Fig. 3d, Au/TiO2–OV presents a photocatalytic NH3 production activity of 78.6 μmol h−1 g−1, ca. 98 and 35 times higher than those of Au/TiO2 and TiO2–OV, respectively. No noticeable decrease in the photocatalytic NH3 generation rate of Au/TiO2–OV is observed over five cycles of testing, suggesting its good stability. Fig. 3e illustrates the photocatalytic NH3 production mechanism of the Au/TiO2–OV system. Under light illumination, hot electrons are produced in Au nanospheres and injected into the TiO2 NS conduction band. Then the hot electrons are trapped in defect states caused by the OV in TiO2 NSs. Subsequently, the hot electrons reduce the N2 molecule adsorption and activation at the OV sites. In the meantime, the hot holes in the Au nanospheres are mainly used up by the methanol as an electron donor. This highly efficient “working-in-tandem” photocatalytic mechanism leads to the outstanding NH3 production performance in the Au/TiO2–OV system. Furthermore, a chemical bath deposition (CBD) approach was adopted by Xiao et al.43 to load Au nanoparticles (NPs) on the surface of (BiO)2CO3 nanodisks (NDs). An improved photocatalytic NH3 generation activity of 38.23 μmol mg−1 h−1 was observed on the as-prepared Au NP deposited (BiO)2CO3 NDs in comparison to that of (BiO)2CO3 NDs alone. This is because Au NPs can increase light absorption and generate hot electrons for N2 reduction as well as rapidly accepting the photo-excited electrons and promoting electron–hole separation.


image file: d0ta11201a-f3.tif
Fig. 3 (a) TEM result of Au/TiO2–OV. (b) EPR spectra for Au/TiO2–OV, Au/TiO2, TiO2–OV and TiO2. (c) N2 TPD profiles for Au/TiO2–OV, Au/TiO2, TiO2–OV and TiO2. (d) Photocatalytic NH3 production activities of various samples using visible light irradiation (λ > 420 nm). (e) Schematic illustration for the production of plasmonic hot electrons in Au NSs, electron transfer into the TiO2 conduction band, and N2 reduction in photocatalytic NH3 production on Au/TiO2–OV using visible-light illumination (λ > 420 nm). Reproduced from ref. 42 with permission from the American Chemical Society. (f) SEM and (g) AFM images of 0.05BPCNS. (h) Average photocatalytic NH3 production activities of various samples in 4 hour reaction. (i) Time course of photocatalytic NH3 production on various samples. (j) Time-resolved photoluminescence spectra of CNS and 0.05BPCNS. (k) Schematic illustration of the possible photocatalytic mechanism on 0.05BPCNS. Reproduced from ref. 48 with permission from Elsevier.

Several semiconductor/semiconductor 0D/2D heterojunctions have also been developed for photocatalytic NH3 production.44,45 For example, 0D AgInS2 NP loaded 2D MXene Ti3C2 NSs were fabricated via in situ growth utilizing a hydrothermal method.44 The as-fabricated 0D/2D AgInS2/Ti3C2 heterostructure displays a photocatalytic NH3 production activity of 38.8 μmol g−1 h−1 utilizing visible light (λ > 400 nm) in 20% methanol aqueous solution. An AQE of 0.07% at 420 nm was also achieved on this 0D/2D AgInS2/Ti3C2 heterostructure. The origin of the excellent photocatalytic NH3-production activity was studied by both experimental characterization and theoretical calculations. The 2D morphology of Ti3C2 NSs and the Z scheme heterostructure formed between AgInS2 and Ti3C2 apparently benefit the photo-induced electron–hole pair dissociation as well as movement. The high specific surface area of Ti3C2 NSs with ample surface reactive sites facilitates the adsorption/activation/reduction of N2. Furthermore, the DFT based theoretical computations also corroborate the spontaneous activation of N2 molecules on Ti3C2 NSs via a di-nuclear end-on bound structure. In addition, Gao et al.45 fabricated a p–n heterojunction of AgCl/δ-Bi2O3 ultrathin NSs by a two-step approach of hydrothermal reaction and precipitation. The AFM characterization indicates a uniform thickness of ca. 2.7 nm for the AgCl/δ-Bi2O3 NSs. Both the high-resolution XPS spectra of O 1s and EPR spectrum confirm the presence of oxygen vacancies. The as-prepared AgCl/δ-Bi2O3 NSs present a photocatalytic NH3 generation activity of 606 μmol h−1 g−1 utilizing visible light (λ > 400 nm), apparently higher than that of δ-Bi2O3 alone. The boosted photocatalytic activity originates from the construction of a p–n heterojunction suppressing the electron–hole recombination and the oxygen vacancies promoting the chemical adsorption/activation of N2. Besides, the ultrathin 2D NS morphology also facilitates the migration of electrons and holes, thus contributing to the improved photocatalytic activity.

Wang et al.46 synthesized three types of polyoxometalates (POMs)/reduced GO composites as semiconductor/semimetal 0D/2D photocatalysts for photocatalytic N2 reduction without any co-catalysts or electron donors. The largest photocatalytic NH3 production activity of 130.3 μmol L−1 h−1 was obtained on H5[PMo10V2O40]/reduced graphene oxide. They attribute the excellent performance to three reasons: (i) reduced aggregation of GO with more exposed reactive sites leading to enhanced N2 adsorption; (ii) broad light absorption range and good reduction capacity; and (iii) GO boosting electron migration and inhibiting electron–hole recombination.

4.2 2D/2D binary heterostructures

The integration of two different 2D building blocks to establish a 2D/2D heterostructure is deemed to be an effective strategy because their intimate contact with large interfacial area greatly favours the dissociation and migration of photo-induced electrons and holes as well as strong interaction for achieving excellent stability.47,48 For instance, Xiao et al.47 have synthesized a 2D/2D MoO2/BiOCl composite via electrostatic adsorption. The as-prepared MoO2/BiOCl heterostructure displays an obviously improved photocatalytic activity of 35 μmol g−1 h−1 for N2 reduction to NH3, compared with either pure MoO2 or bare BiOCl, using xenon light irradiation. They ascribed the improved activity to the presence of MoO2 which not only promoted charge carrier migration but also facilitated the adsorption/activation of N2. Furthermore, Qiu et al.48 combined black phosphorus (BP) NSs with g-C3N4 NSs to form 0.05BPCNS via calcination in an Ar atmosphere. The SEM image of 0.05BPCNS displays a rough surface due to the loading of BP NSs (Fig. 3f). The thickness of 0.05BPCNS is shown to be ca. 4–5 nm in its AFM image (Fig. 3g). Furthermore, as displayed in Fig. 3h, 0.05BPCNS presents an averaged photocatalytic NH3 generation activity of 347.5 μmol L−1 h−1, much larger than that of g-C3N4 NSs (40.5 μmol L−1 h−1) or BP NSs (45.3 μmol L−1 h−1). The time course of photocatalytic NH3 production of 0.05BPCNS in Fig. 3i confirms its excellent robustness. The XPS characterization results indicate the formation of C–P covalent bonds between g-C3N4 NSs and BP NSs, bringing about improved charge carrier separation/transfer efficiency and increased stability owing to the occupation of lone electron pairs on the P atom. This is also corroborated by the time-resolved photoluminescence spectra (Fig. 3j), which indicate the longer charge carrier lifetime of 0.05BPCNS compared with that of CNS. The photocatalytic mechanism in Fig. 3k indicates that the photo-induced electrons on the CNS conduction band are transported to BPNS, where the N2 molecules are adsorbed and reduced to form NH3. In the meantime, the photo-induced holes in the CNS valence band oxidize the methanol.

4.3 3D/2D binary heterostructures

Feng et al.49 fabricated a LnCO3OH (Ln = La, Pr) coupled g-C3N4 heterostructure using a hydrothermal approach. Fig. 4a shows the TEM image of the LaCO3OH–CN composite. A hexagonal morphology is observed. Moreover, PrCO3OH with an irregular morphology is dispersed on CN (Fig. 4b). As shown in Fig. 4c, LaCO3OH/g-C3N4 and PrCO3OH/g-C3N4 exhibit improved photocatalytic NH3 generation activities of 8200 and 8900 μmol h−1 g−1, respectively. Their activities are much larger than that of g-C3N4 or hydrothermally treated g-C3N4. Besides, both LaCO3OH/g-C3N4 and PrCO3OH/g-C3N4 display robust stabilities over five cycles of reaction. The XRD and high-resolution XPS spectral results of the used LaCO3OH/g-C3N4 and PrCO3OH/g-C3N4 also confirm the good stabilities of their crystal and chemical structures. Furthermore, the results of the EPR spin-trap test with DMPO in water (Fig. 4d) show four characteristic peaks with an intensity ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]1 on LaCO3OH/g-C3N4, ascribed to DMPO–˙OH. Thus, a Z-scheme heterojunction is proposed to form in LaCO3OH/g-C3N4 (Fig. 4e). First, N2 is adsorbed onto the photocatalysts surface. Upon light illumination, the photo-excited electrons in the LaCO3OH conduction band recombine with the photo-excited holes in the g-C3N4 valence band. The photo-excited electrons on the g-C3N4 conduction band reduce the adsorbed N2 to generate NH3. Simultaneously, the photo-excited holes on the LaCO3OH valence band oxidize the OH to form ˙OH. The formed ˙OH is then captured by methanol as the scavenger. Therefore, they ascribe the excellent photocatalytic NH3-production performance of LnCO3OH/g-C3N4 to the strong chemical adsorption of N2 and the efficient electron–hole separation caused by the Z-scheme heterojunction with the Ln–N electron migration channel.
image file: d0ta11201a-f4.tif
Fig. 4 TEM results of (a) La–CN and (b) Pr–CN. (c) Photocatalytic NH3 production activities of CN, HCN, LaCO3OH, PrCO3OH, La–CN and Pr–CN. (d) The EPR spectra of DMPO–˙OH in the presence of La–CN. (e) The photocatalytic NH3 production mechanism of La–CN. Reproduced from ref. 49 with permission from Elsevier. (f) TEM image of TiO2@C/g-C3N4. (g) Photocatalytic NH3 production activities of various samples using visible light illumination (λ > 420 nm). (h) Schematic illustration for the band alignment and photo-induced electron–hole pair dissociation and migration in the TiO2@C/g-C3N4 system. Reproduced from ref. 51 with permission from RSC.

4.4 Ternary heterostructures

2D building blocks have also been applied in complex ternary heterostructured photocatalysts to achieve synergistically enhanced photocatalytic NH3 generation.50,51 For example, Xing et al.50 fabricated a ternary photocatalyst of MoS2 NP loaded carbon coated ZnO NSs (MoS2/C–ZnO) via combining hydrothermal and photo-deposition approaches. The optimized 1 wt% MoS2 loaded carbon coated ZnO (1% MoS2/C–ZnO) shows the largest photocatalytic NH3 production activity of 245.7 μmol L−1 g−1 h−1 using simulated sunlight illumination. This activity is 9.3 and 4.0 times larger than that of ZnO and carbon coated ZnO (C–ZnO), respectively. They ascribed the high activity on 1% MoS2/C–ZnO to the boosted dissociation/migration of charge carriers due to the presence of a carbon layer and MoS2 as electron trappers. The enlarged surface area also contributes to the improved activity. In contrast, C–ZnO exhibits the largest photocatalytic NH3 production activity under visible-light irradiation, since the mixed MXene Ti3C2Tx and melamine was calcined to yield 2D carbon nanosheet-supported TiO2 NPs wrapped with g-C3N4 NSs (TiO2@C/g-C3N4).51 The TEM image of TiO2@C/g-C3N4 (Fig. 4f) indicates that TiO2 NPs are dispersed onto the surfaces of C NSs and coupled with g-C3N4 NSs. The high-resolution XPS spectrum of Ti 2p confirms the presence of a large amount of Ti3+ in TiO2@C/g-C3N4. As presented in Fig. 4g, an excellent visible-light-driven photocatalytic NH3 generation rate of 250.6 μmol h−1 g−1 accompanied by a quantum yield of 0.14% at 420 nm was obtained on the as-prepared TiO2@C/g-C3N4. This activity is not only ca. 18 and 10 times higher than that of g-C3N4 and TiO2@C, respectively, but also larger than that of a physically mixed TiO2@C and g-C3N4 sample (TiO2@C + g-C3N4). The authors attribute the outstanding activity of TiO2@C/g-C3N4 to its ample surface defects, strong electron-donating capacity, suitable light absorption capability, efficient charge migration and outstanding N2 activation capacity. Besides, the type II heterostructure established between TiO2 and g-C3N4 also improves the efficient electron–hole dissociation and transportation (Fig. 4h).

For the 2D material based heterostructured photocatalysts discussed in Sections 4.1–4.4, it is of central importance to design and fabricate appropriate heterojunctions (e.g., type II heterojunction, p–n junction and Z scheme) with efficient interfacial charge transfer and a strongly bonded interface to accomplish high activity and stability in photocatalytic NH3 production. Besides, both reduction and oxidation active sites could be accommodated on different components in the heterojunction on the basis of the migration direction of photo-induced electrons and holes. In particular, the unique advantages of 2D materials, e.g., high surface area, ultrathin thickness and abundant active sites, are greatly beneficial for achieving the above target.

5. Conclusions and outlook

In summary, the design and preparation of single-component and heterostructured photocatalysts using two-dimensional (2D) building blocks are summarized and introduced. The comprehensive and insightful composition/structure–performance relationships in these 2D material based photocatalysts for ammonia production are discussed. The precise and in-depth reaction mechanisms of these 2D material based photocatalysts are also elucidated.

Although some achievements have been made in the above area, there are still many approaches to be developed. For example, it is of great importance to rationally design and fabricate single-component 2D photocatalysts with the following properties: (i) appropriate band gap width for a broad light-responsive range; (ii) desired band edge positions for sufficient redox ability of photo-induced electrons and holes towards the N2 reduction reaction and oxidation reaction (e.g., water oxidation); (iii) ultrathin thickness and high crystallinity for efficient electron–hole dissociation and transportation; (iv) large surface area and massive active sites (e.g., cation/anion vacancies, doped heteroatoms and specific crystal facets with abundant undercoordinated surface atoms). Apart from the four intriguing above-mentioned properties, the rational design and synthesis of heterostructured 2D material based photocatalysts with the following characteristics is essential: (i) complementary band gap widths of different components in the heterostructures for maximizing the light absorption range; (ii) favourable band alignment (e.g., type II heterojunction, p–n junction and Z scheme heterojunction) for high-efficiency electron–hole separation and transfer; (iii) compatible crystal structures of different components for achieving strong binding, thus facilitating fast interfacial charge migration and high stability; (iv) creation of reduction and oxidation active sites on different components based on the electron–hole transfer direction in the heterojunction.

Furthermore, both state-of-the-art characterization techniques, e.g., aberration-corrected scanning transmission electron microscopy, synchrotron radiation-based X-ray absorption spectroscopy, in situ Fourier transform infrared spectroscopy and in situ Raman, and powerful theoretical calculations can be combined together to explore the overall and in-depth composition/structure–performance correlation in 2D material based photocatalysts. Besides, a range of newly developed in situ or operando characterization methods, e.g., in situ transmission electron microscopy, in situ electron spin resonance spectroscopy, in situ Raman and operando synchrotron-based X-ray absorption spectroscopy, can be adopted to reveal the actual reaction mechanisms in the photocatalytic ammonia production procedure. The acquired insightful and overall composition–structure/performance relationship, together with the revealed reaction mechanisms will further contribute to the design and preparation of brand-new high-performance photocatalysts utilizing 2D building blocks towards light-driven ammonia generation. Moreover, it is highly promising to adopt powerful theoretical computations for a high-throughput screening of photocatalysts with novel compositions and structures. Then these predicted photocatalysts can be rationally designed and fabricated using advanced nanotechnology to achieve efficient, highly selective and steady light-induced ammonia production.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The authors gratefully acknowledge financial support from the Australian Research Council (ARC) through the Discovery Project programs (FL170100154, DP160104866, DP170104464 and DE200100629) and the Linkage Project program (LP160100927).

References

  1. C. Guo, J. Ran, A. Vasileff and S.-Z. Qiao, Energy Environ. Sci., 2018, 11, 45 RSC.
  2. J. Li, H. Li, G. Zhan and L. Zhang, Acc. Chem. Res., 2017, 50, 112 CrossRef CAS.
  3. Z. Wang, X. Hu, Z. Liu, G. Zou, G. Wang and K. Zhang, ACS Catal., 2019, 9, 10260 CrossRef CAS.
  4. X. Chen, N. Li, Z. Kong, W.-J. Ong and X. Zhao, Mater. Horiz., 2018, 5, 9 RSC.
  5. K. Ithisuphalap, H. Zhang, L. Guo, Q. Yang, H. Yang and G. Wu, Small Methods, 2019, 3, 1800352 CrossRef.
  6. C. Ling, Y. Ouyang, Q. Li, X. Bai, X. Mao, A. Du and J. Wang, Small Methods, 2019, 3, 1800376 CrossRef.
  7. X. Yan, D. Liu, H. Cao, F. Hou, J. Liang and S. X. Dou, Small Methods, 2019, 3, 1800501 CrossRef.
  8. M. Li, H. Huang, J. Low, C. Gao, R. Long and Y. Xiong, Small Methods, 2019, 3, 1800388 CrossRef.
  9. A. J. Medford and M. C. Hatzell, ACS Catal., 2017, 7, 2624 CrossRef CAS.
  10. D. Yan, H. Li, C. Chen, Y. Zou and S. Wang, Small Methods, 2019, 3, 1800331 CrossRef.
  11. Y. Zhao, R. Shi, X. Bian, C. Zhou, Y. Zhao, S. Zhang, F. Wu, G. I. N. Waterhouse, L.-Z. Wu, C.-H. Tung and T. Zhang, Adv. Sci., 2019, 6, 1802109 CrossRef.
  12. C. Mao, J. Wang, Y. Zou, H. Li, G. Zhan, J. Li, J. Zhao and L. Zhang, Green Chem., 2019, 21, 2852 RSC.
  13. M. Cheng, C. Xiao and Y. Xie, J. Mater. Chem. A, 2019, 7, 19616 RSC.
  14. Q. Wang, J. Guo and P. Chen, J. Energy Chem., 2019, 36, 25 CrossRef.
  15. M.-H. Vu, M. Sakar, S. A. Hassanzadeh-Tabrizi and T.-O. Do, Adv. Mater. Interfaces, 2019, 6, 1900091 CrossRef.
  16. X. Xue, R. Chen, C. Yan, P. Zhao, Y. Hu, W. Zhang, S. Yang and Z. Jin, Nano Res., 2019, 12, 1229 CrossRef CAS.
  17. G. Y. Duan, Y. Ren, Y. Tang, Y. Z. Sun, Y. M. Chen, P. Y. Wan and X. J. Yang, ChemSusChem, 2020, 13, 88 CrossRef CAS.
  18. R. Li, Chin. J. Catal., 2018, 39, 1180 CrossRef CAS.
  19. H. Kisch, Eur. J. Inorg. Chem., 2020, 1376 CrossRef CAS.
  20. G. Zhang, C. D. Sewell, P. Zhang, H. Mi and Z. Lin, Nano Energy, 2020, 71, 104645 CrossRef CAS.
  21. Q. Hao, C. Liu, G. Jia, Y. Wang, H. Arandiyan, W. Wei and B.-J. Ni, Mater. Horiz., 2020, 7, 1014 RSC.
  22. J. S. J. Hargreavesa, Y.-M. Chung, W.-S. Ahnc, T. Hisatomid, K. Domen, M. C. Kung and H. H. Kung, Appl. Catal., A, 2020, 594, 117419 CrossRef.
  23. Y. Huang, N. Zhang, Z. Wu and X. Xie, J. Mater. Chem. A, 2020, 8, 4978 RSC.
  24. Y. Xue, Y. Guo, Z. Liang, H. Cui and J. Tian, J. Colloid Interface Sci., 2019, 556, 206 CrossRef CAS.
  25. Y. Zhang, J. Di, P. Ding, J. Zhao, K. Gu, X. Chen, C. Yan, S. Yin, J. Xia and H. Li, J. Colloid Interface Sci., 2019, 553, 530 CrossRef CAS.
  26. W. Wang, H. Zhang, S. Zhang, Y. Liu, G. Wang, C. Sun and H. Zhao, Angew. Chem., Int. Ed., 2019, 58, 16644 CrossRef CAS.
  27. S. Cao, B. Fan, Y. Feng, H. Chen, F. Jiang and X. Wang, Chem. Eng. J., 2018, 353, 147 CrossRef CAS.
  28. X. Lv, W. Wei, F. Li, B. Huang and Y. Dai, Nano Lett., 2019, 19, 6391 CrossRef CAS.
  29. C. Ren, Y. Zhang, Y. Li, Y. Zhang, S. Huang, W. Lin and K. Ding, J. Phys. Chem. C, 2019, 123, 17296 CrossRef CAS.
  30. D. Wu, R. Wang, C. Yang, Y. An, H. Lu, H. Wang, K. Cao, Z. Gao, W. Zhang, F. Xu and K. Jiang, J. Colloid Interface Sci., 2019, 556, 111 CrossRef CAS.
  31. J. Di, J. Xia, M. F. Chisholm, J. Zhong, C. Chen, X. Cao, F. Dong, Z. Chi, H. Chen, Y.-X. Weng, J. Xiong, S.-Z. Yang, H. Li, Z. Liu and S. Dai, Adv. Mater., 2019, 31, 1807576 CrossRef.
  32. X. Xue, R. Chen, H. Chen, Y. Hu, Q. Ding, Z. Liu, L. Ma, G. Zhu, W. Zhang, Q. Yu, J. Liu, J. Ma and Z. Jin, Nano Lett., 2018, 18, 7372 CrossRef CAS.
  33. Y. Zhao, Y. Zhao, G. I. N. Waterhouse, L. Zheng, X. Cao, F. Teng, L.-Z. Wu, C.-H. Tung, D. O'Hare and T. Zhang, Adv. Mater., 2017, 29, 1703828 CrossRef.
  34. Y. Zhao, L. Zheng, R. Shi, S. Zhang, X. Bian, F. Wu, X. Cao, G. I. N. Waterhouse and T. Zhang, Adv. Energy Mater., 2020, 2002199 CrossRef CAS.
  35. S. Zhang, Y. Zhao, R. Shi, C. Zhou, G. I. N. Waterhouse, L.-Z. Wu, C.-H. Tung and T. Zhang, Adv. Energy Mater., 2020, 10, 1901973 CrossRef CAS.
  36. Y. Zhao, Y. Zhao, R. Shi, B. Wang, G. I. N. Waterhouse, L.-Z. Wu, C.-H. Tung and T. Zhang, Adv. Mater., 2019, 31, 1806482 CrossRef.
  37. Y. Li, X. Chen, M. Zhang, Y. Zhu, W. Ren, Z. Mei, M. Gu and F. Pan, Catal. Sci. Technol., 2019, 9, 803 RSC.
  38. S. Sun, X. Li, W. Wang, L. Zhang and X. Sun, Appl. Catal., B, 2017, 200, 323 CrossRef CAS.
  39. T. Hou, R. Guo, L. Chen, Y. Xie, J. Guo, W. Zhang, X. Zheng, W. Zhu, X. Tan and L. Wang, Nano Energy, 2019, 65, 104003 CrossRef CAS.
  40. S. Liu, Y. Wang, S. Wang, M. You, S. Hong, T.-S. Wu, Y.-L. Soo, Z. Zhao, G. Jiang, J. Qiu, B. Wang and Z. Sun, ACS Sustainable Chem. Eng., 2019, 7, 6813 CrossRef CAS.
  41. P. Huang, W. Liu, Z. He, C. Xiao, T. Yao, Y. Zou, C. Wang, Z. Qi, W. Tong, B. Pan, S. Wei and Y. Xie, Sci. China: Chem., 2018, 61, 1187 CrossRef CAS.
  42. J. Yang, Y. Guo, R. Jiang, F. Qin, H. Zhang, W. Lu, J. Wang and J. C. Yu, J. Am. Chem. Soc., 2018, 140, 8497 CrossRef CAS.
  43. C. Xiao, H. Hu, X. Zhang and D. R. MacFarlane, ACS Sustainable Chem. Eng., 2017, 5, 10858 CrossRef CAS.
  44. J. Qin, X. Hu, X. Li, Z. Yin, B. Liu and K. Lam, Nano Energy, 2019, 61, 27 CrossRef CAS.
  45. X. Gao, Y. Shang, L. Liu and F. Fu, J. Catal., 2019, 371, 71 CrossRef CAS.
  46. X.-H. Li, W.-L. Chen, H.-Q. Tan, F.-R. Li, J.-P. Li, Y.-G. Li and E.-B. Wang, ACS Appl. Mater. Interfaces, 2019, 11, 37927 CrossRef CAS.
  47. C. Xiao, H. Wang, L. Zhang, S. Sun and W. Wang, ChemCatChem, 2019, 11, 6467 CrossRef CAS.
  48. P. Qiu, C. Xu, N. Zhou, H. Chen and F. Jiang, Appl. Catal., B, 2018, 221, 27 CrossRef CAS.
  49. X. Feng, H. Chen, F. Jiang and X. Wang, Chem. Eng. J., 2018, 347, 849 CrossRef CAS.
  50. P. Xing, P. Chen, Z. Chen, X. Hu, H. Lin, Y. Wu, L. Zhao and Y. He, ACS Sustainable Chem. Eng., 2018, 6, 14866 CrossRef CAS.
  51. Q. Liu, L. Ai and J. Jiang, J. Mater. Chem. A, 2018, 6, 4102 RSC.
  52. J. Ran, H. Zhang, J. Qu, J. Shan, S. Chen, F. Yang, R. Zheng, J. Cairney, L. Song, L. Jing and S.-Z. Qiao, ACS Mater. Lett., 2020, 2, 1484 CrossRef CAS.
  53. G. Fiori, F. Bonaccorso, G. Iannaccone, T. Palacios, D. Neumaier, A. Seabaugh, S. K. Banerjee and L. Colombo, Nat. Nanotechnol., 2014, 9, 768 CrossRef CAS.
  54. J. Ran, J. Qu, H. Zhang, T. Wen, H. Wang, S. Chen, L. Song, X. Zhang, L. Jing, R. Zheng and S.-Z. Qiao, Adv. Energy Mater., 2019, 9, 1803402 CrossRef.
  55. J. Ran, H. Wang, H. Jin, C. Ling, X. Zhang, H. Ju, L. Jing, J. Wang, R. Zheng and S.-Z. Qiao, J. Mater. Chem. A, 2018, 6, 23278 RSC.
  56. J. Ran, B. Zhu and S.-Z. Qiao, Angew. Chem., Int. Ed., 2017, 56, 10373 CrossRef CAS.
  57. T. Carey, S. Cacovich, G. Divitini, J. Ren, A. Mansouri, J. M. Kim, C. Wang, C. Ducati, R. Sordan and F. Torrisi, Nat. Commun., 2017, 8, 1202 CrossRef.
  58. B. Xia, J. Ran, S. Chen, L. Song, X. Zhang, L. Jing and S.-Z. Qiao, Nanoscale, 2019, 11, 8304 RSC.
  59. J. Ran, G. Gao, F.-T. Li, T.-Y. Ma, A. Du and S.-Z. Qiao, Nat. Commun., 2017, 8, 13907 CrossRef CAS.
  60. J. Ran, X. Wang, B. Zhu and S.-Z. Qiao, Chem. Commun., 2017, 53, 9882 RSC.
  61. K. F. Mak and J. Shan, Nat. Photonics, 2016, 10, 216 CrossRef CAS.
  62. J. Ran, T. Y. Ma, G. Gao, X.-W. Du and S. Z. Qiao, Energy Environ. Sci., 2015, 8, 3708 RSC.
  63. J. Ran, W. Guo, H. Wang, B. Zhu, J. Yu and S.-Z. Qiao, Adv. Mater., 2018, 30, 1800128 CrossRef.

This journal is © The Royal Society of Chemistry 2021