Next Article in Journal
Synthesis and Spectroscopic Properties of Sm3+-Activated Li6Y(BO3)3 Phosphor for Light-Emitting Diode Applications
Previous Article in Journal
A Mathematical Model for Predicting the Ultra-Early-Age Strength of Concrete
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of Metals and Their Compounds/Black Phosphorus-Based Nanomaterials in the Direction of Photocatalytic Hydrogen Evolution

1
Basic Experimental Center for Natural Science, University of Science and Technology Beijing, Beijing 100083, China
2
School of Chemistry and Biological Engineering, University of Science and Technology Beijing, Beijing 100083, China
3
Hubei Three Gorges Laboratory, Yichang 443007, China
*
Author to whom correspondence should be addressed.
Coatings 2024, 14(9), 1141; https://doi.org/10.3390/coatings14091141
Submission received: 15 July 2024 / Revised: 31 August 2024 / Accepted: 3 September 2024 / Published: 5 September 2024

Abstract

:
Black phosphorous (BP) is a novel composite material. Its carrier mobility can reach more than 1000 cm2·V−1·s−1 and has a direct bandgap adjustable from 0.3 to 1.5 eV with thickness, so its photovoltaic performance is good. These properties show great potential for applications in many fields, such as energy storage, sensors, biomedicine, and environmental treatment. With the deepening of research, it is found that the instability of BP under natural environmental conditions and the limitations of its preparation limit its development, while combining with other materials can further optimize its performance, which not only improves the mechanical properties of the material but also gives it new functions. Based on this, this paper summarizes the preparation and optical properties of highly stable metals and their compounds/BP-based nanomaterials in recent years, highlights the progress of their application in photocatalytic hydrogen evolution, and gives an outlook on the challenges and opportunities for the future development of BP in photocatalysis.

1. Introduction

On 1 March 2024, the International Energy Agency (IEA) released its Global Carbon Emissions 2023 report. The report points out that global energy-related carbon dioxide emissions reached a record 37.4 billion tons in 2023, an increase of 410 million tons, or 1.1%, compared with the previous year [1], which will make mankind face more serious environmental problems and global climate change [2]. Therefore, the Peak Carbon Neutrality Program not only calls for reducing the use of fossil energy and vigorously promoting clean energy but also calls for repairing the environmental problems that have already been caused [3,4].
Photocatalytic technology, including photocatalytic hydrogen evolution, photocatalytic reduction of CO2, and artificial nitrogen fixation, provides a new technological pathway for clean hydrogen production, low-energy-consumption nitrogen fixation, and CO2 reuse [5]. Therefore, photocatalytic technology is of great significance in realizing carbon neutrality and the carbon cycle.
BP materials, as an emerging two-dimensional (2D) material, have received great attention from researchers in recent years. BP crystals consist of layers of lamellar structures stacked one on top of another, and the layers are bonded to each other by van der Waals forces so that they can be exfoliated to form a 2D structure [6,7]. Compared with other two-dimensional materials, BP materials have the following advantages: (1) the bandgap of BP is tunable and decreases with the increase of layer thickness, the forbidden band width of BP in bulk phase is 0.3 eV, while that of monolayer varies up to 1.5 eV; (2) the absorption range is wide, covering the UV-visible-near-infrared region so that solar energy can be utilized by as much as 95%; (3) The electron mobility of BP can reach up to 1000 cm2·V−1·s−1, which is beneficial for carriers to rapidly migrate to the material surface, promoting the separation of electron-hole pairs and participating in subsequent redox reactions, therefore enhancing the catalytic performance. (4) 2D BP material has a larger specific surface area and more abundant active sites, which is conducive to the adsorption on the surface of the reactants and the occurrence of the redox reaction. The surface adsorption of reactants and redox reactions are favored [8]. Based on the above characteristics, BP materials are the most promising materials, which have been widely studied and applied in the fields of photo catalysis and environmental treatment.
However, despite the great potential of BP nanosheets, their stability, biocompatibility, and large-scale preparation remain challenges in current research [9,10,11,12]. To overcome these problems, researchers have explored different synthesis methods, surface modification techniques, and composite designs and found that BP can be further optimized by forming heterojunctions or composites with other materials. This is mainly reflected in the following two points: (1) The creation of heterostructures can broaden the range of spectral absorption, enhances the ability of light absorption, and increases the density of photogenerated carriers; (2) The formation of heterojunctions enhances the separation and migration of photogenerated carriers, therefore expediting the dissociation of electron-hole pairs. This process subsequently augments the stability of BP materials. Therefore, this paper summarizes the preparation and optical properties of high-stability BP-based nanomaterials in recent years, focuses on the research progress of their application in photocatalytic hydrogen evolution, and presents a forward-looking perspective on the challenges and opportunities that lie ahead for the future development of BP-based nanocomposites within the realm of photocatalysis.

2. Properties of BP Materials

2.1. Crystal Structure of BP

Monomorphic phosphorus has a variety of isoforms, commonly red, white, and black phosphorus. Among them, black phosphorus is more stable and does not spontaneously combust or ignite easily when exposed to air [13]. BP crystals consist of a single layer of BP stacked on top of each other. In each BP monolayer, phosphorus atoms are bonded to three other atoms, three of which are in the same plane, and the fourth atom is in a parallel neighboring plane, which finally forms a honeycomb structure, as shown in Figure 1 [14]. The BP layers are connected through BP layers are connected by weak van der Waals forces, and the spacing between the layers is between 3.21 and 3.73 Å. Therefore, BP nanomaterials with smaller sizes can be obtained by exfoliation of bulk crystals, and then, according to the number of layers and sizes, BPs can be categorized into quantum dots, monolayers, oligo layers, and blocks [15,16].

2.2. Bandgap Modulation of BP

BPs are direct bandgap semiconductors from monolayers to blocks, but the forbidden bandwidth of BPs is affected by the number of layers; the fewer the layers, the greater the forbidden bandwidth, up to a span of 0.3 eV to 2.0 eV [17]. In 2014, Cai [18] et al. investigated in detail the relationship between the number of BP layers and the forbidden bandwidth by first-nature theory calculations. Figure 2 illustrates the variation in bandgap with thickness, ranging from 1 L to 5 L, as well as the trends observed in the valence and conduction bands. Consequently, it can be inferred that the bandgap, band alignment, and carrier effective mass are all contingent upon the number of layers. Given that a lower carrier mass typically indicates higher mobility, it follows that the bandgap, band alignment, figure of merit, and carrier effective mass are all influenced by the number of layers. The performance of few-layer phosphorene may potentially surpass that of single-layer phosphorene, therefore offering a plethora of new opportunities for a wide array of electronic applications.
Besides the number of layers, doping, construction of heterojunction, composite, and external electric field [19,20,21] can also regulate the forbidden bandwidth of BP. Researchers can obtain materials with different bandgaps by adjusting the thickness, which in turn can obtain materials with different light absorption properties and expand the light absorption region. The improvements are mainly reflected in the following three points: (1) The formation of the composite material helps to synthesize the light absorption performance of the material, broaden the light absorption range and enhance the light absorption, and increase the concentration of photogenerated carriers; (2) It is conducive to the improvement of the photogenerated carrier separation rate of catalysts, accelerating the spatial separation of electron-hole pairs, and at the same time, the migration and reaction consumption of carriers contribute to the improvement of the stability of the BP; (3) The composites can form close interactions at the interface, which on the one hand facilitates the effective separation of carriers at the interface and on the other hand enhances the stability of BP.

2.3. Stability of the BP

Since BP is a layered material with a graphene-like structure, its lattice consists of bi-atomic layers, each made up of zigzagging chains of phosphorus atoms. This structure allows black phosphorus to be chemically or physically exfoliated into nanosheets of varying thickness. The theoretical maximum specific surface area of a single layer of black phosphorus has been calculated to be up to 2400 m2/g, which can provide many active sites for catalytic reactions and thus facilitate the catalytic process. However, BP itself is unstable and easily oxidized. Figure 3 provides an atomic force microscopy (AFM) image of a BP nanoflake (~9.0 nm thick) shortly after exfoliation, and it was found that the bumps became progressively larger and taller with increasing environmental exposure, suggesting that the BP nanoflake was continuously oxidizing [22]. Figure 4 shows the use of X-ray photoelectron spectroscopy (XPS) to assess whether chemical modification of BP occurs under environmental exposure, and the exfoliated BP is consistent with previous XPS measurements of black phosphorus bulk crystals at 0 h of environmental exposure. After 13 h of environmental exposure, the full width of the half peak of black phosphorus increased, and the peak height increased with time, which is consistent with the oxidation state of phosphorus (POx), indicating that BP is easily degraded in the natural environment. BP was first photo-oxidized to form PxOy and then absorbed water to form vesicles on the surface of 2D-BP, which became larger and larger with time and then formed phosphoric acid or phosphite, which was finally lost from BP. The vesicles gradually become larger with time and finally form phosphoric acid or phosphite from BP.
Researchers have investigated the mechanism of BP oxidation in the environment from various perspectives, including experiments and theories. The results showed that the easily oxidizable property of BP is related to the lone pair of free electrons of phosphorus atoms and that the incompletely bonded edge phosphorus atoms may produce unstable binding structures [23]. In 2015, Ziletti et al. [24] showed through first-principles calculations that the BP surface is susceptible to degradation by reacting with oxygen. Favron et al. [25] conducted experiments that demonstrated the oxidation of BP in response to oxygen, water, and visible light. The rate of this oxidation was found to be influenced by both the concentration of oxygen and the intensity of light, as depicted in Figure 5. Therefore, effective passivation of BP and improvement of the environmental stability of BP are of great significance for the construction of BP nanodevices and application development. At present, domestic and foreign researchers have used a variety of means to improve the stability of 2D-BP in the environment, such as covalent or noncovalent functionalization modification on the surface of BP or the surface adsorption of metal cations, etc., and the introduction of protective layers such as graphene.

2.4. Optical Properties of BP

BP not only has a unique electronic structure but also exhibits a series of compelling optical properties. These characteristics offer significant potential for its application in optoelectronics, sensing, and energy conversion fields. Woomer et al. [26] employed the liquid-phase exfoliation technique to derive monolayers, bilayers, or few-layered 2D-BP nanosheets and subsequently investigated their optical properties. The findings are depicted in Figure 6. We can see that as the thickness of the sheet decreases, the spectrum is gradually blue-shifted, and its enhanced optoelectronic properties, mechanical properties, and chemical reaction activity are also increased. In 2019, Huang et al. [27], using infrared spectroscopy, demonstrated that in-plane biaxial strain can effectively modulate the van der Waals interactions between few-layered BPs with thicknesses from 2 to 10 layers and significantly enhance electron transport along the serrated direction of the BPs. The calculation results show that the optical response of BP can be significantly modulated by strain engineering. In 2016, Hersam et al. [28] used a liquid-phase exfoliation method to prepare and obtain few-layer 2D-BP nanosheets, and their optical properties were characterized, and the optical absorption spectra showed that there were two peaks between 900 and 1250 nm [29,30]. From these studies, it can be seen that the optical properties of BP are affected by several factors, such as the number of layers, structure, strain, and doping. First, there is a significant difference in the optical properties between the monolayer and multilayer structures of BP, which can be further modulated by changing the structure of BP, such as twisting or applying strain, to achieve a fine control of its function. Second, the interfaces between BP and other materials also have an impact on its optical properties, thus changing its electronic structure and optical response. Furthermore, the optical properties of BP are modulated by doping or introducing other elements. These changes in optical properties will provide new opportunities for the application of BP in fields such as optoelectronic devices, biosensing, and photonics.

3. Metals and Their Compounds/BP-Based Nanomaterials

3.1. Characterization of Metals and Their Compounds/BP-Based Nanomaterials

Bonding in BP is characterized by robust intralayer P-P bonds and relatively weak interlayer van der Waals forces, and its structure is similar to that of graphite. These robust P-P covalent bonds give rise to a nonplanar bilayer structure oriented in the sawtooth direction [31]. Two-dimensional BP materials exhibit a range of properties, including a tunable bandgap, high carrier mobility, large specific surface area, and superior optoelectronic performance, which make them show excellent potential for biomedical, drug delivery, and photocatalysis applications [32,33]. The surface of BP nanomaterials is very prone to oxidation, forming phosphates. However, phosphate groups have a strong binding ability to metal elements [34]. Therefore, 2D BP nanosheets and BP quantum dots are also successfully combined with metal nanomaterials to construct metal/BP-based nanocomposites. Metal/BP nanocomposites not only retain the inherent properties of BP but also incorporate several distinctive characteristics of metal-based nanomaterials, including imaging contrast ability and catalytic effects. The advancement of these nanocomposites has demonstrated potential applications in the field of photocatalysis [35].

3.2. Preparation of Metals and Their Compounds/BP-Based Nanomaterials

In recent years, researchers have investigated a multitude of techniques for synthesizing metal/BP-based nanocomposites that require combining metal or metal-based nanomaterials with BP nanomaterials. In this article, the preparation of BP nanomaterials and how metal or metal-based nanomaterials can be constructed into heterojunctions with BP nanomaterials will be emphasized.
Usually, the preparation methods of BP nanomaterials (BP nanosheets and BP quantum dots) mainly include mechanical exfoliation, chemical vapor deposition, liquid-phase stripping method, electrochemical method, etc. In the case of 2D BP nanosheets, the interaction between the two layers is predominantly governed by a weak van der Waals force so that they can be separated by the mechanical exfoliation method, but the morphology is not well controlled, and the yield is low. The liquid-phase stripping method offers an effective solution to this issue, characterized by its simplicity and cost-effectiveness. Within the liquid-phase stripping process, organic solvents such as N-methyl-2-pyrrolidone, N, N-dimethylformamide, and dimethyl sulfoxide are frequently employed to disperse BP nanosheets within an aqueous solution. Nevertheless, the stability of BP nanomaterials in aqueous solution is low and prone to aggregation and degradation. To solve this problem, deoxidized water surfactant was added to the dispersing solution, which resulted in good dispersion of BP in aqueous solvent [36].
Both the mechanical stripping method and the liquid-phase stripping method have certain limitations in controlling the morphology of the generated BP nanosheets. In contrast, the chemical vapor deposition method can regulate the thickness of the BP nanosheets, thus generating better morphology and larger BP nanosheets. Compared with the preparation of BP nanosheets, BP quantum dots are mainly prepared by the mechanical stripping method and liquid-phase stripping method.
The main methods for constructing metal/BP-based nano complexes by combining metal or metal-based nanomaterials with BP nanomaterials are doping and in situ growth methods. Through electrostatic interaction, BP can affix most metal or metal-based nanomaterials of opposing charges, therefore forming metal/BP-based nanocomposites. Xu [37] et al. coated upconversion nanoparticles (UCNPs) on mesoporous silica, then electrostatically attracted copper sulfide (CuS) nanoparticles with strong negative charges, and finally combined them with a monolayer of BP nanosheets, the preparation process is shown in Figure 7a. However, the manipulation of metal/BP-based nanomaterial morphology and the loading capacity of BP nanomaterials via simple electrostatic adsorption presents significant challenges. At this time, the in situ growth method offers an effective solution to the aforementioned issues. Wu [38] et al. modified ultra-thin BP nanosheets with PEG (PEG-BPN) and used this as a template to deposit MnO2 nanosheets on top. Then, BPN/MnO2 nanomaterials were prepared at room temperature using the in situ self-sacrificial reduction method of KMnO4. The preparation process is shown in Figure 7b.
Although some metal/BP-based nanocomposites have been successfully constructed using the above preparation methods, relatively few studies have been conducted. Further investigations into more streamlined and adaptable techniques are warranted to enhance their efficiency and versatility.

4. Application of Metals and Their Compounds/BP-Based Nanomaterials in Photocatalytic Hydrogen Evolution

Hydrogen is an important renewable and clean energy source, and in 1972, it was first discovered that light TiO2 electrodes can crack water to produce hydrogen, so the use of solar energy to produce hydrogen cannot only effectively alleviate the energy problem but also green technology [39,40,41,42,43,44,45].
The reaction equation for cracking aqueous hydrogen is as follows:
H2O = 1/2O2 + H2 △G = +237 KJ/mol (Endothermic reaction)
The use of semiconductors for photocatalysis requires that their forbidden bandwidth should be greater than 1.23 eV and the lowest position of the conduction band should be below the reduction potential of hydrogen [46]. Only then can H+ be reduced to produce H2 [47,48]. Zhu [49] et al. reported that two-dimensional BP nanosheets have certain hydrogen production properties under alkaline conditions, but the hydrogen production performance is low, and the stability is poor. It was possible to improve the material stability by modifying the surface of BP, but its hydrogen production performance was still low [50,51,52,53]. Therefore, the introduction of metals and their compounds, etc., to form heterojunctions in complex with BP can improve the effective separation of photogenerated electron-hole pairs, further enhance the photocatalytic hydrogen production performance on the one hand, and broaden the scope of utilization of sunlight on the other hand, as well as improve the stability of the BP materials.

4.1. Photocatalytic Hydrogen Evolution Properties of Metal-BP Composites

Precious metal materials such as platinum (Pt) [54], silver (Ag) [55], and gold (Au) [56] have excellent electrical conductivity, which facilitates the fast transfer of photogenerated carriers in semiconducting materials and thus improves the separation efficiency of electron-hole pairs. In 2018, Tian et al. [57] synthesized BP nanosheets by solvothermal method at 60–140 °C as shown in Figure 8a, and constructed heterojunctions by loading Pt on BP nanosheets, and the hydrogen production performance increased by about 30-fold under visible light irradiation as shown in Figure 8b. To make the co-catalyst Pt more stably loaded on the carrier surface, Xue [58] et al. grew Pt nanoparticles in situ on the surface of BP nanosheets, where BP nanosheets provide a larger surface area and abundant exposed active sites for the confined growth of highly dispersed Pt nanoparticles, thus enhancing the interaction between the interface of BP nanosheets and Pt, which leads to the effective separation of electron-hole pair which enhances the absorption of sunlight and greatly improves the hydrogen production of 89.1 mmol after a 6 h reaction production.
Overall, the loading of noble metals on BP can enhance photocatalysis for the following three main reasons: first, the localized surface plasmon resonance (LSPR) effect of noble metals can increase the light absorption intensity of the composites in the visible region; second, under visible light irradiation, the phenomenon of localized electric field enhancement occurs at the boundaries of noble metal particles, which induces the photogenerated charges generated at the near-surface of the BP to be rapidly transferred to the noble metal nanoparticles; finally, the photogenerated holes are retained on the valence band of BP, which also has the ability of photocatalysis. Therefore, the noble metal-catalyzed hydrogen production ability of BP can be significantly improved by taking advantage of the LSPR effect of the noble metal as well as the high electrical conductivity.

4.2. Photocatalytic Hydrogen Evolution Performance of Metal Oxide and BP Composites

Compositing other metal compounds with 2D BP nanosheets to form heterojunctions, as shown in Figure 9a, can effectively solve the problem of easy carrier compounding while enhancing the cyclic stability of BP [59]. The heterojunctions formed by metal compounds/BP composites include type I, type II, type Z, type S, and multicomponent heterojunction photocatalysts, where multicomponent heterogeneity refers to the inclusion of two or more heterogeneous structures in the reaction system [60,61,62,63]. TiO2 is the most widely studied metal oxide, and it is also a chemically stable and low-cost semiconductor material. It can be utilized for photocatalysis using sunlight, but the efficiency is low. To solve this problem, it can be compounded with spectrally responsive materials to form heterojunctions, thus broadening the light absorption range and accelerating the spatial separation of carriers [64]. Elbanna [65] et al. prepared BP/TMC composites by compositing TiO2 mesocrystals (TMC) with BP, which greatly improved the solar light responsivity, resulting in an increase in the hydrogen production rate of the obtained composites under both visible (λ > 420 nm) and near-infrared light (λ > 780 nm). From the valence band mapping, the conduction band energy levels of BP and TMC are located at 1.06 eV and 0.45 eV, respectively, and the valence band energy levels are located at −0.23 eV and −2.75 eV, respectively, and the composite energy band structure conforms to the type II heterojunction, as shown in Figure 9b.
Bismuth-based materials are also promising in the field of photocatalytic hydrogen production due to their good visible light responsiveness [66,67,68]. Hu et al. [69] modified Bi2WO6 on the surface of BP nanosheets using the hydrothermal method and self-assembly technique to form layer-stacked composites. Due to the excellent light absorption properties of the BP materials, the intensity of the absorption peak at 400–800 nm of the composites was enhanced with the increase of the black phosphorus content, and the visible hydrogen production performance was optimal when the black phosphorus loading reached 12 wt.%, which was about 21,042 μmol·g−1. The carrier migration paths under the light conditions are shown in Figure 10a. Zhu [70] et al. prepared BP/BiVO4 composites on this basis to realize photocatalytic water evolution hydrogen reaction. From Figure 10b, It can be seen that the hydrogen evolution mechanism is similar to that of BP/Bi2WO6, and the reasonable construction of Z-type heterojunction can improve the carrier composite problem of the photocatalytic material and enhance its redox ability at the same time. In summary, the construction of heterostructures can significantly improve the efficiency of photocatalytic water splitting. The combination of 2D materials such as BP and bismuth vanadate provides strong support for achieving this goal while also opening up new possibilities for the development of novel efficient photocatalysts. Future research may continue to explore more combinations of 2D materials and optimization strategies to further improve photocatalytic performance and stability.

4.3. Photocatalytic Hydrogen Evolution Properties of Metal Sulfide and BP Composites

Sulfides also have suitable forbidden bandwidths and have attracted much attention in the field of visible light photocatalytic hydrogen production. Ran et al. [71] complexed two-dimensional BP nanosheets with different sulfides (CdS, ZnS) to form photocatalysts to investigate the improvement of sulfide photocatalytic performance by BP nanosheets with different thicknesses. It has been shown that few-layer BP nanosheets (about 5–6 layers and 4–5 nm thick) can enhance the hydrogen production rate of sulfide, and when 1 wt.% of few-layer BP nanosheets are loaded on the surface of CdS, it can not only enhance the stability of sulfide, but also improve its visible light hydrogen production performance, and the rate of hydrogen production can be up to 11,192 μmol·h−1·g−1. Based on this, Ran et al. [72] loaded the 0D ZnxCd1−xS (ZCS) nanoparticles loaded on few-layer phosphorene (FLP). X-ray near-edge absorption spectroscopy and UV photoelectron spectroscopy results show that the composite interface has interactions so that the photogenerated electrons-hole pairs are effectively separated at the interface, which improves the hydrogen production rate of the material under visible light conditions. The hydrogen production mechanism is shown in Figure 11a.
MoS2 is considered to be an ideal semiconductor photocatalyst due to its excellent hydrogen production performance. It has a layered structure similar to that of BP materials and can form 2D/2D contact with BP, which can shorten the charge migration distance and thus improve hydrogen production activity. Yuan et al. [73] constructed a 2D/2D heterojunction composite BP/MoS2 with a hydrogen generation rate of up to 1286 μmol·h−1·g−1 under visible light conditions. In addition, it was found in cycling experiments that the photocatalytic performance of the composite was stable, with a decrease of only 8%. After the combination of BP and MoS2, the contact surfaces between BP and MoS2 are very close, which can shorten the charge transport distance, which is conducive to the rapid separation of photogenerated carriers, and then under the light conditions, the photogenerated electrons on BP will migrate to the surface of MoS2 and react with the adsorbed H+ to generate H2. Zhu et al. [74] complexed WS2 with BP and achieved a 21-fold increase in the hydrogen production rate in the NIR region compared to BP alone because WS2 has a lower figure of merit (5.57 eV), which favors photogenerated charge transfer and hydrogen evolution reactions in the black phosphorus conduction zone, the hydrogen production mechanism is shown in Figure 11b. To further broaden the response range of absorption spectra, Reddy et al. [75] synthesized composite nanorods of BP/MoS2/1D CdS by a hydrothermal method, and the ternary composites showed a hydrogen production rate of up to 183.24 mmol−1·g−1 in the visible light with the help of the excellent light absorption ability and abundant active sites of CdS nanorods, the hydrogen production mechanism is shown in Figure 11c. Mao et al. [76] developed a ternary heterostructure consisting of BP quantum dots, CdS nanoparticles, and LTO nanostrata as an effective noble-metal-free photocatalyst, and the BP-CdS-LTO complexes under solar irradiation produced hydrogen at a rate of 0.96 mmol·g−1·h−1. Zhang et al. [77] showed that other sulfides, such as ZnIn2S4, were also used to improve the catalytic performance of BP materials. Based on an in-depth understanding of charge transfer between semiconductors with different energy bands, this study opens a new avenue for a more rational design of heterostructure photocatalysts for hydrogen production.
Figure 11. Schematic diagram of the photocatalytic H2 generation mechanism under visible light irradiation. (a) 0D/2D ZCS/ (FLP) [72]. (b) BP/WS2 in the presence of EDTA [74]. (c) CdS/BP-MoS2 [75].
Figure 11. Schematic diagram of the photocatalytic H2 generation mechanism under visible light irradiation. (a) 0D/2D ZCS/ (FLP) [72]. (b) BP/WS2 in the presence of EDTA [74]. (c) CdS/BP-MoS2 [75].
Coatings 14 01141 g011

4.4. Photocatalytic Hydrogen Evolution Performance of BP-Based Ternary Heterojunction Materials

Zhu [78] and others constructed a ternary BP/Pt/RGO composite photocatalytic material using reduced graphene oxide (RGO) as a charge transfer channel, Pt as a co-catalyst, and BP nanosheets as a catalyst, and the rates of hydrogen production were 5.13 μmol in the visible (λ > 420 nm) and 1.26 μmol in the near-infrared (λ > 780 nm), respectively. The photocurrent density was enhanced, and resistance decreased by introducing RGO, indicating that the interfacial charge transfer rate was accelerated to transfer electrons from BP to Pt via RGO rapidly, and then the reduction reaction with H+ was carried out to generate H2, improving the production rate. After the introduction of RGO, the photocurrent density is enhanced, and the resistance is reduced, which indicates that the interfacial charge transfer rate is accelerated to transfer the electrons from BP to Pt via RGO, and then the reduction reaction with H+ occurs to generate H2, which improves the rate of hydrogen production. Zhu [79] et al. used Au nanoparticles as a “binder” to form a BP/Au/La2Ti2O7 ternary heterojunction, which has good hydrogen production performance in the near-infrared (NIR) region at λ > 780 nm, with a hydrogen production rate of 0.30 mmol·g−1·h−1.

5. Conclusions

In this paper, the application of BP in photocatalysis is introduced in detail, and the comparison of hydrogen production performance is shown in Table 1. In photocatalytic applications, BP shows many unique advantages, such as rich morphology, a large number of active sites, a wide range of light absorption, etc., as well as good adaptability to the combination of BP and other materials, which makes BP play multiple roles in photocatalysis. In addition to the above advantages, BP also exhibits some problems, such as its easy oxidization, which has a great influence on the cyclic stability of photocatalysis and makes it more sensitive to the reaction environment. In addition, the low yield of BP nanosheets and quantum dots is the main reason hindering the rapid application of BP and wider research. Although the mechanism of BP in photocatalysis is well understood, deeper studies are still needed to optimize the BP materials in a more targeted way to improve their catalytic performance and stability.

6. Prospects

By means of compositing or constructing heterojunctions of 2D BP materials with metals, metal oxides, and metal sulfides, the separation efficiency of photogenerated electron-hole pairs can be effectively improved to enhance the light-absorbing properties of the materials on the one hand, and the stability of the photocatalysts can be improved on the other hand. The rich variety of metals and their compounds can provide more choices for BP-based composites.
In summary, it can be inferred that the future research directions of BP are as follows:
(1)
During the synthesis of BP nanosheets, the following problems still exist: the size and thickness of the nanosheets are not controllable, they are easy to oxidize, the yield is low, and the production cost is high. It is expected that in the future, we can realize the large-scale production of two-dimensional BP nanosheets with accurately controllable thickness and size through a new preparation method or a combination of multiple preparation methods.
(2)
The preparation of black phosphorus-based materials can be optimized by means of in situ testing and DFT calculations; however, there are still fewer studies in this area, and DFT calculations can be further promoted to optimize the design of high-performance black phosphorus-based materials. In addition, the structural changes in photocatalytic materials and their microscopic kinetic processes in the catalytic reaction are deeply investigated, and the in-depth photocatalytic mechanism study provides ideas for designing a reasonable catalyst structure.
(3)
The photocatalytic active sites of nano-BP are still unclear, and the mechanism of photocatalytic hydrogen evolution needs to be further explored and elaborated.
This review summarizes and analyzes the research on the combination of metals and their compounds with BP in the field of photocatalytic hydrogen evolution in recent years, hoping to provide some new ideas for researchers and scholars and some theoretical foundations for the early development of low-cost, high-efficiency and stable photocatalysts.

Author Contributions

Conceptualization, W.Z. and L.Q.; software, W.Z. and B.Y.; formal analysis, W.Z. and H.Y.; investigation, W.Z. and X.L.; writing—original draft preparation, W.Z.; writing—review and editing, W.Z.; supervision, L.Q.; funding acquisition, W.Z. and S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Open/Innovation Fund of Hubei Three Gorges Laboratory, grant number SK232005, and College Students’ Innovative Entrepreneurial Training Plan Program, grant number 202499202A, and Teaching Reform Project of Beijing University of Science and Technology, grant number JG2022M52.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. IEA. CO2 Emissions in 2023. Available online: https://www.iea.org/reports/co2-emissions-in-2023 (accessed on 5 March 2024).
  2. Liu, Z.; Deng, Z.; Davis, S.J.; Ciais, P. Global carbon emissions in 2023. Nat. Rev. Earth Environ. 2024, 5, 253–254. [Google Scholar] [CrossRef]
  3. Chien, A.A. Driving the cloud to true zero carbon. Commun. ACM 2021, 64, 5. [Google Scholar] [CrossRef]
  4. Chen, W.D.; Han, M.Z.; Bi, J.Y.; Meng, Y. China can peak its energy-related CO2 emissions before 2030: Evidence from driving factors. J. Clean. Prod. 2023, 429, 139584. [Google Scholar] [CrossRef]
  5. Wang, G.J.; Lv, S.S.; Shen, Y.H.; Li, W.; Lin, L.H.; Li, Z.C. Advancements in heterojunction, cocatalyst, defect and morphology engineering of semiconductor oxide photocatalysts. J. Mater. 2024, 10, 335–338. [Google Scholar] [CrossRef]
  6. Thomas, S.A. Layered two-dimensional black phosphorous-based hybrid electrodes for rechargeable batteries. J. Energy Storage 2023, 73, 109068. [Google Scholar] [CrossRef]
  7. Tian, Y.; Wang, H.D.; Li, H.N.; Guo, Z.N.; Tian, B.N.; Cui, Y.X.; Li, Z.F.; Li, G.H.; Zhang, H.; Wu, Y.C. Recent advances in black phosphorus/carbon hybrid composites: From improved stability to applications. J. Mater. Chem. 2020, 8, 4647–4676. [Google Scholar] [CrossRef]
  8. Pang, X.; Xue, S.X.; Zhou, T.; Yuan, H.D.; Liu, C.; Lei, W.Y. Advances in Two-Dimensional Black Phosphorus-Based Nanostructures for Photocatalytic Applications. Prog. Chem. 2022, 34, 630–642. [Google Scholar]
  9. Zhang, X.Y.; Li, Z.; Wang, Y.; Zhang, S.H.; Zang, X.H.; Wang, C.; Wang, Z. Preparation of black phosphorus nanosheets/zeolitic imidazolate framework nanocomposite for high-performance solid-phase microextraction of organophosphorus pesticides. J. Chromatogr. A 2023, 1708, 464339. [Google Scholar] [CrossRef]
  10. Wang, R.; Zhang, M.; Ge, B.C.; Zhang, L.X.; Zhou, J.X.; Liu, S.F.; Jiao, T.F. Facile preparation of black phosphorus-based rGO-BP-Pd composite hydrogels with enhanced catalytic reduction of 4-nitrophenol performances for wastewater treatment. J. Mol. Liq. 2020, 310, 113083. [Google Scholar] [CrossRef]
  11. Li, H.; Liu, A.M.; Ren, X.F.; Yang, Y.N.; Gao, L.G.; Fan, M.Q.; Ma, T.L. A black phosphorus/Ti3C2 MXene nanocomposite for sodium-ion batteries: A combined experimental and theoretical study. Nanoscale 2019, 11, 19862. [Google Scholar] [CrossRef]
  12. Li, C.; Cui, X.M.; Gao, X.; Liu, S.Z.; Sun, Q.; Lian, H.Q.; Zu, L.; Liu, Y.; Wang, X.D.; Cui, X.G. Electrochemically prepared black phosphorene micro-powder as flame retardant for epoxy resin. Compos. Interfaces 2021, 28, 693–705. [Google Scholar] [CrossRef]
  13. Nilges, T.; Kersting, M.; Pfeifer, T. A fast low-pressure transport route to large black phosphorus single crystals. J. Solid State Chem. 2008, 181, 1707–1711. [Google Scholar] [CrossRef]
  14. Chaudhary, V.; Neugebauer, P.; Mounkachi, O.; Lahbabi, S.; Fatimy, A.E. Phosphorene-an emerging two-dimensional material: Recent advances in synthesis, functionalization, and applications. 2D Mater. 2022, 9, 032001. [Google Scholar] [CrossRef]
  15. Kou, L.Z.; Chen, C.F.; Smith, S.C. Phosphorene: Fabrication, Properties, and Applications. J. Phys. Chem. Lett. 2015, 6, 2794–2805. [Google Scholar] [CrossRef] [PubMed]
  16. Batmunkh, M.; Bat-Erdene, M.; Shapter, J.G. Phosphorene and Phosphorene-Based Materials—Prospects for Future Applications. Adv. Mater. 2016, 28, 8586–8617. [Google Scholar] [CrossRef]
  17. Jakobczyk, P.; Kaczmarzyk, I.; Bogdanowicz, R. Few-Layer Black Phosphorus/Chitosan Nanocomposite Electrodes via Controlled Electrodeposition for Enhanced Electrochemical Kinetic Performance. J. Phys. Chem. C 2024, 128, 7087–7095. [Google Scholar] [CrossRef]
  18. Cai, Y.Q.; Zhang, G.; Zhang, Y.W. Layer-dependent Band Alignment and Work Function of Few-Layer Phosphorene. Sci. Rep. 2014, 4, 6677. [Google Scholar] [CrossRef] [PubMed]
  19. Hanlon, D.; Backes, C.; Doherty, E.; Cucinotta, C.S.; Berner, N.C.; Boland, C.; Lee, K.; Harvey, A.; Lynch, P.; Gholamvand, Z.; et al. Liquid exfoliation of solvent-stabilized few-layer black phosphorus for applications beyond electronics. Nat. Commun. 2015, 6, 8563. [Google Scholar] [CrossRef]
  20. Zhao, W.C.; Xue, Z.M.; Wang, J.F.; Jiang, J.Y.; Zhao, X.H.; Mu, T.C. Large-Scale, Highly Efficient, and Green Liquid-Exfoliation of Black Phosphorus in Ionic Liquids. ACS Appl. Mater. Intertfaces 2015, 7, 27608–27612. [Google Scholar] [CrossRef]
  21. Shen, Z.K.; Yuan, Y.J.; Yu, Z.T.; Zou, Z.G. Progress on the Black Phosphorus Nanomaterials for Photocatalysis. Mater. China 2021, 40, 493–507. [Google Scholar]
  22. Wood, J.D.; Wells, S.A.; Jariwala, D.; Chen, K.S.; Cho, E.K.Y.; Sangwan, V.K.; Liu, X.L.; Lauhon, L.J.; Marks, T.J.; Hersam, M.C. Effective Passivation of Exfoliated Black Phosphorus Transistors against Ambient Degradation. Nano Lett. 2014, 14, 6964–6970. [Google Scholar] [CrossRef] [PubMed]
  23. Luo, W.; Zemlyanov, D.Y.; Milligan, C.A.; Du, Y.C.; Yang, L.M.; Wu, Y.Q.; Ye, P.D. Surface chemistry of black phosphorus under a controlled oxidative environment. Nanotechnology 2016, 27, 434002. [Google Scholar] [CrossRef]
  24. Ziletti, A.; Carvalho, A.; Campbell, D.K.; Coker, D.F.; Neto, A.H.C. Oxygen Defects in Phosphorene. Phys. Rev. Lett. 2015, 114, 046801. [Google Scholar] [CrossRef]
  25. Favron, A.; Gaufrès, E.; Fossard, F.; Phaneuf-L’Heureux, A.L.; Tang, N.Y.W.; Lévesque, P.L.; Loiseau, A.; Leonelli, R.; Francoeur, S.; Martel, R. Photooxidation and quantum confinement effects in exfoliated black phosphorus. Nat. Mater. 2015, 14, 826–832. [Google Scholar] [CrossRef]
  26. Woomer, A.H.; Farnsworth, T.W.; Hu, J.; Wells, R.A.; Donley, C.L.; Warren, S.C. Phosphorene: Synthesis, Scale-Up, and Quantitative Optical Spectroscopy. ACS Nano 2015, 9, 8869–8884. [Google Scholar] [CrossRef] [PubMed]
  27. Huang, S.Y.; Zhang, G.W.; Fan, F.R.; Song, C.Y.; Wang, F.J.; Xing, Q.X.; Wang, C.; Wu, H.; Yan, H.G. Strain-tunable van der Waals interactions in few-layer black phosphorus. Nat. Commun. 2019, 10, 2447. [Google Scholar] [CrossRef]
  28. Kang, J.; Wells, S.A.; Wood, J.D.; Lee, J.H.; Liu, X.L.; Ryder, C.R.; Zhu, J.; Guest, J.R.; Husko, C.A.; Hersam, M.C. Stable aqueous dispersions of optically and electronically active phosphorene. Proc. Natl. Acad. Sci. USA 2016, 113, 11688–11693. [Google Scholar] [CrossRef] [PubMed]
  29. Yasaei, P.; Kumar, B.; Foroozan, T.; Wang, C.H.; Asadi, M.; Tuschel, D.; Indacochea, J.E.; Klie, R.F.; Salehi-Khojin, A. High quality black phosphorus atomic layers by liquid-phase exfoliation. Adv. Mater. 2015, 27, 1887–1892. [Google Scholar] [CrossRef]
  30. Huang, W.C.; Zi, Y.; Wang, M.K.; Hu, L.P.; Wang, Y.Z.; Xie, Z.J.; Qiu, M.; Zhang, H. Advancements in photophysics research and applications of phosphorene semiconductors. J. Shenzhen Univ. Sci. Eng. 2024, 41, 323–347. [Google Scholar] [CrossRef]
  31. Choi, J.R.; Yong, K.W.; Choi, J.Y.; Nilghaz, A.; Lin, Y.; Xu, J.; Lu, X. Black Phosphorus and its Biomedical Applications. Theranostics 2018, 8, 1005–1026. [Google Scholar] [CrossRef]
  32. Lei, W.Y.; Liu, G.; Zhang, J.; Liu, M.H. Black phosphorus nanostructures: Recent advances in hybridization, doping and functionalization. Chem. Soc. Rev. 2017, 46, 3492–3509. [Google Scholar] [CrossRef] [PubMed]
  33. Miao, Y.L.; Shi, X.T.; Li, Q.T.; Hao, L.J.; Liu, L.; Liu, X.; Chen, Y.H.; Wang, Y.J. Engineering natural matrices with black phosphorus nanosheets to generate multi-functional therapeutic nanocomposite hydrogels. Biomater. Sci. 2019, 7, 4046–4059. [Google Scholar] [CrossRef] [PubMed]
  34. He, X.L.; Wu, J.X.; Huang, X.L.; Chen, Y.; Zhang, L.; Sheng, X.X. Three-in-one polymer nanocomposite coating via constructing tannic acid functionalized MXene/BP hybrids with superior corrosion resistance, friction resistance, and flame-retardancy. Chem. Eng. Sci. 2024, 283, 119429. [Google Scholar] [CrossRef]
  35. Wang, J.W.; Qiu, S.L.; Chen, W.J.; Yang, W.H.; Song, L.; Hu, Y. Synthesis of Nitrogen-Rich Black Phosphorus via a Diazotization Strategy for Efficient Flame Retardancy and Mechanical Properties of Epoxy Resins. Macromol. Mater. Eng. 2022, 307, 2200143. [Google Scholar] [CrossRef]
  36. Liu, Y.; Wang, Y.H.; Song, S.Y.; Zhang, H.J. Tumor Diagnosis and Therapy Mediated by Metal Phosphorus-Based Nanomaterials. Adv. Mater. 2021, 33, 2103936. [Google Scholar] [CrossRef] [PubMed]
  37. Xu, M.S.; Yang, G.X.; Bi, H.T.; Xu, J.T.; Dong, S.M.; Jia, T.; Wang, Z.; Zhao, R.X.; Sun, Q.Q.; Gai, S.L. An intelligent nanoplatform for imaging-guided photodynamic/photothermal/chemo-therapy based on upconversion nanoparticles and CuS integrated black phosphorus. Chem. Eng. J. 2020, 382, 122822. [Google Scholar] [CrossRef]
  38. Wu, Q.; Chen, G.; Gong, K.K.; Wang, J.; Ge, X.X.; Liu, X.Q.; Guo, S.J.; Wang, F. MnO2-Laden Black Phosphorus for MRI-Guided Synergistic PDT, PTT, and Chemotherapy. Matter 2019, 1, 496–512. [Google Scholar] [CrossRef]
  39. Gao, R.R.; Cheng, B.; Fan, J.J.; Yu, J.G.; Ho, W.K. ZnxCd1-xS quantum dot with enhanced photocatalytic H2-production performance. Chin. J. Catal. 2021, 42, 15–24. [Google Scholar] [CrossRef]
  40. Yang, F.; Ai, Y.; Li, X.; Wang, L.; Zhang, Z.; Ding, W.; Sun, W. Preparation of electrochemical horseradish peroxidase biosensor with black phosphorene-zinc oxide nanocomposite and their applications. RSC Adv. 2023, 13, 32028–32038. [Google Scholar] [CrossRef]
  41. Drmosh, Q.A.; Alade, I.O.; Alkanad, K.; Alnaggar, G.; Khan, A.; Khan, M.Y.; Onaizi, S.A. Fabrication of Z-scheme TiO2//BP/g-C3N4 nanocomposite via pulsed laser ablation in liquid for photocatalytic overall water splitting. Opt. Mater. 2022, 128, 112428. [Google Scholar] [CrossRef]
  42. Li, S.; Zhang, F.; Wang, J.; Wen, W.; Wang, S. Black phosphorus-Au nanocomposite-based fluorescence immunochromatographic sensor for high-sensitive detection of zearalenone in cereals. Nanophotonics 2020, 9, 2397–2406. [Google Scholar] [CrossRef]
  43. Li, X.; Wang, L.; Yan, L.; Han, X.; Zhang, Z.; Zhang, X.; Sun, W. A Portable Wireless Intelligent Nanosensor for 6,7-Dihydroxycoumarin Analysis with A Black Phosphorene and Nano-Diamond Nanocomposite-Modified Electrode. Biosensors 2023, 13, 153. [Google Scholar] [CrossRef] [PubMed]
  44. Zhou, Y.; Qiu, S.; Chu, F.; Liu, W.; Yang, W.; Hu, W.; Hu, Y. Passivation of black phosphorus by triazine-based silica coating: Hierarchical BP@SiO2-N@Co(OH)2 structure for enhanced fire safety and toughness of unsaturated polyester resins. Compos. Part A-Appl. Sci. Manuf. 2023, 164, 107279. [Google Scholar] [CrossRef]
  45. Zhang, J.; Lu, X.; Shi, X. Facile synthesis of amorphous carbon-coated bismuth phosphate nanocomposite for lithium-ion battery anode with ultra-long cycle stability. J. Solid State Chem. 2022, 316, 123588. [Google Scholar] [CrossRef]
  46. Du, W.; Chen, W.; Wang, J.; Zhang, H.; Song, L.; Hu, Y.; Ma, X. A dual-nanozyme-loaded black phosphorus multifunctional therapeutic platform for combined photothermal/photodynamic/starvation cancer therapy. J. Mater. Chem. B 2023, 11, 5185–5194. [Google Scholar] [CrossRef] [PubMed]
  47. Joshi, G.; Saha, A.; Dutta, A.; Khatua, S. NIR-Driven Photocatalytic Hydrogen Production by Silane- and Tertiary Amine-Bound Plasmonic Gold Nanoprisms. ACS Appl. Mater. Interfaces 2022, 14, 38815–38823. [Google Scholar] [CrossRef]
  48. Savateev, O.; Shvalagin, V.; Tang, J. Increasing Profitability of Ethanol Photoreforming by Simultaneous Production of H2 and Acetal. Glob. Chall. 2024, 8, 2300078. [Google Scholar] [CrossRef]
  49. Zhu, H.; Trinh, M.T.; Wang, J.; Fu, Y.; Joshi, P.P.; Miyata, K.; Zhu, X.Y. Organic Cations Might Not Be Essential to the Remarkable Properties of Band Edge Carriers in Lead Halide Perovskites. Adv. Mater. 2017, 29, 1603072. [Google Scholar] [CrossRef]
  50. Lu, B.C.; Zheng, X.Y.; Li, Z.S. A promising photocatalyst for water-splitting reactions with a stable sandwiched P4O2/black phosphorus heterostructure and high solar-to-hydrogen efficiency. Nanoscale 2020, 12, 6617–6623. [Google Scholar] [CrossRef]
  51. Vishnoi, P.; Gupta, U.; Pandey, R.; Rao, C.N.R. Stable functionalized phosphorenes with photocatalytic HER activity. J. Mater. Chem. A 2019, 7, 6631–6637. [Google Scholar] [CrossRef]
  52. Ren, J.; Yu, Y.; Qu, Y.N.; Sun, D.F.; Chu, W.H.; Su, Q.M.; Huang, L.Y.; Xu, B.S. Bionic construction of a dual Z-scheme MoO3/ZnIn2S4/black phosphorus quantum dots heterojunction with expanded redox surfaces and photoelectron transfer ability for high-efficiency photocatalysis. Chem. Eng. J. 2023, 465, 142894. [Google Scholar] [CrossRef]
  53. Shi, R.; Liu, F.; Wang, Z.; Weng, Y.; Chen, Y. Black/red phosphorus quantum dots for photocatalytic water splitting: From a type I heterostructure to a Z-scheme system. Chem. Commun. 2019, 55, 12531–12534. [Google Scholar] [CrossRef]
  54. Ma, Z.; Li, P.; Ye, L.; Wang, L.; Xie, H.; Zhou, Y. Selectivity reversal of photocatalytic CO2 reduction by Pt loading. Catal. Sci. Technol. 2018, 8, 5129–5132. [Google Scholar] [CrossRef]
  55. Teixeira, G.F.; Silva Junior, E.; Vilela, R.; Zaghete, M.A.; Colmati, F. Perovskite Structure Associated with Precious Metals: Influence on Heterogenous Catalytic Process. Catalysts 2019, 9, 721. [Google Scholar] [CrossRef]
  56. Zhang, L.; Liang, Q.; Ning, H.; Yang, P.; Yan, J. Photocatalytic Performance of CuO/CuAl2O4 Composite Photocatalysts Modified by Precious Metals (Pt,Ag,Ru). Min. Metall. Eng. 2018, 38, 115–118,122. [Google Scholar]
  57. Tian, B.; Tian, B.; Smith, B.; Scott, M.C.; Lei, Q.; Hua, R.; Liu, Y. Facile bottom-up synthesis of partially oxidized black phosphorus nanosheets as metal-free photocatalyst for hydrogen evolution. Proc. Natl. Acad. Sci. USA 2018, 115, 4345–4350. [Google Scholar] [CrossRef] [PubMed]
  58. Xue, Y.; Min, S.; Wang, F. Dye-sensitized black phosphorus nanosheets decorated with Pt cocatalyst for highly efficient photocatalytic hydrogen evolution under visible light. Int. J. Hydrogen Energy 2019, 44, 21873–21881. [Google Scholar] [CrossRef]
  59. Shen, Z.K.; Yuan, Y.J.; Pei, L.; Yu, Z.T.; Zou, Z. Black phosphorus photocatalysts for photocatalytic H2 generation: A review. Chem. Eng. J. 2020, 386, 123997. [Google Scholar] [CrossRef]
  60. Lim, S.K.; Kang, S.C.; Yoo, T.J.; Lee, S.K.; Hwang, H.J.; Lee, B.H. Operation Mechanism of a MoS2/BP Heterojunction FET. Nanomaterials 2018, 8, 797. [Google Scholar] [CrossRef]
  61. Wang, G.C.; Wu, L.M.; Yan, J.H.; Zhou, Z.; Ma, R.S.; Yang, H.F.; Li, J.J.; Gu, C.Z.; Bao, L.H.; Du, S.X.; et al. Intrinsic charge transport behaviors in graphene-black phosphorus van der Waals heterojunction devices. Chin. Phys. B 2018, 27, 077303. [Google Scholar] [CrossRef]
  62. Wang, J.; Liu, F.; Li, Y.; Chen, L.; Chen, Y.; Zhang, H.; Xie, Z. Enhancement of Broadband Reverse Saturable Absorption of Red/Black Phosphorus Heterojunction. Molecules 2024, 29, 1271. [Google Scholar] [CrossRef] [PubMed]
  63. Zhang, T.; Zhou, L.; Chen, G.; Xia, S.; Qiu, M. Theoretical study on an oxygen-modified phosphorene autogenous Z-scheme heterojunction for hydrogen evolution. Chem. Commun. 2023, 59, 1517–1520. [Google Scholar] [CrossRef]
  64. Guo, Q.; Zhou, C.Y.; Ma, Z.B.; Yang, X.M. Fundamentals of TiO2 Photocatalysis: Concepts, Mechanisms, and Challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef] [PubMed]
  65. Elbanna, O.; Zhu, M.; Fujitsuka, M.; Majima, T. Black Phosphorus Sensitized TiO2 Mesocrystal Photocatalyst for Hydrogen Evolution with Visible and Near-Infrared Light Irradiation. ACS Catal. 2019, 9, 3618–3626. [Google Scholar] [CrossRef]
  66. Dutta, V.; Chauhan, A.; Verma, R.; Gopalkrishnan, C.; Nguyen, V.H. Recent trends in Bi-based nanomaterials: Challenges, fabrication, enhancement techniques, and environmental applications. Beilstein J. Nanotechnol. 2022, 13, 1316–1336. [Google Scholar] [CrossRef]
  67. Jiao, Z.; Zhang, Y.; Ouyang, S.; Yu, H.; Lu, G.; Ye, J.; Bi, Y. BiAg Alloy Nanospheres: A New Photocatalyst for H2 Evolution from Water Splitting. ACS Appl. Mater. Interfaces 2014, 6, 19488–19493. [Google Scholar] [CrossRef] [PubMed]
  68. Zhao, L.; Fang, W.; Meng, X.; Wang, L.; Bai, H.; Li, C. In-situ synthesis of metal Bi to improve the stability of oxygen vacancies and enhance the photocatalytic activity of Bi4O5Br2 in H2 evolution. J. Alloys Compd. 2022, 910, 164883. [Google Scholar] [CrossRef]
  69. Hu, J.D.; Chen, D.Y.; Mo, Z.; Li, N.J.; Xu, Q.F.; Li, H.; He, J.H.; Xu, H.L.; Lu, J.M. Z-Scheme 2D/2D Heterojunction of Black Phosphorus/Monolayer Bi2WO6 Nanosheets with Enhanced Photocatalytic Activities. Angew. Chem.-Int. Ed. 2019, 58, 2073–2077. [Google Scholar] [CrossRef]
  70. Zhu, M.; Sun, Z.; Fujitsuka, M.; Majima, T. Z-Scheme Photocatalytic Water Splitting on a 2D Heterostructure of Black Phosphorus/Bismuth Vanadate Using Visible Light. Angew. Chem.-Int. Ed. 2018, 57, 2160–2164. [Google Scholar] [CrossRef]
  71. Ran, J.; Zhu, B.; Qiao, S.Z. Phosphorene Co-catalyst Advancing Highly Efficient Visible-Light Photocatalytic Hydrogen Production. Angew. Chem.-Int. Ed. 2017, 56, 10373–10377. [Google Scholar] [CrossRef]
  72. Ran, J.; Wang, X.; Zhu, B.; Qiao, S.Z. Strongly interactive 0D/2D hetero-structure of a ZnxCd1−xS nano-particle decorated phosphorene nano-sheet for enhanced visible-light photocatalytic H2 production. Chem. Commun. 2017, 53, 9882–9885. [Google Scholar] [CrossRef] [PubMed]
  73. Yuan, Y.J.; Wang, P.; Li, Z.J.; Wu, Y.Z.; Bai, W.F.; Su, Y.B.; Guan, J.; Wu, S.T.; Zhong, J.S.; Yu, Z.T.; et al. The role of bandgap and interface in enhancing photocatalytic H2 generation activity of 2D-2D black phosphorus/MoS2 photocatalyst. Appl. Catal. B-Environ. 2019, 242, 1–8. [Google Scholar] [CrossRef]
  74. Zhu, M.; Zhai, C.; Fujitsuka, M.; Majima, T. Noble metal-free near-infrared-driven photocatalyst for hydrogen production based on 2D hybrid of black Phosphorus/WS2. Appl. Catal. B-Environ. 2018, 221, 645–651. [Google Scholar] [CrossRef]
  75. Reddy, D.A.; Kim, E.H.; Gopannagari, M.; Kim, Y.; Kumar, D.P.; Kim, T.K. Few layered black phosphorus/MoS2 nanohybrid: A promising co-catalyst for solar driven hydrogen evolution. Appl. Catal. B-Environ. 2019, 241, 491–498. [Google Scholar] [CrossRef]
  76. Mao, L.; Cai, X.; Yang, S.; Han, K.; Zhang, J. Black phosphorus-CdS-La2Ti2O7 ternary composite: Effective noble metal-free photocatalyst for full solar spectrum activated H2 production. Appl. Catal. B-Environ. 2019, 242, 441–448. [Google Scholar] [CrossRef]
  77. Zhang, Q.; Zhang, J.; Zhang, L.; Cao, M.; Yang, F.; Dai, W.L. Facile construction of flower-like black phosphorus nanosheet@ZnIn2S4 composite with highly efficient catalytic performance in hydrogen production. Appl. Surf. Sci. 2020, 504, 144366. [Google Scholar] [CrossRef]
  78. Zhu, M.; Osakada, Y.; Kim, S.; Fujitsuka, M.; Majima, T. Black phosphorus: A promising two dimensional visible and near-infrared-activated photocatalyst for hydrogen evolution. Appl. Catal. B-Environ. 2017, 217, 285–292. [Google Scholar] [CrossRef]
  79. Zhu, M.; Cai, X.; Fujitsuka, M.; Zhang, J.; Majima, T. Au/La2Ti2O7 Nanostructures Sensitized with Black Phosphorus for Plasmon-Enhanced Photocatalytic Hydrogen Production in Visible and Near-Infrared Light. Angew. Chem.-Int. Ed. 2017, 56, 2064–2068. [Google Scholar] [CrossRef]
Figure 1. Characterizations of BP. (a) Multilayer BP. (b) Monolayer BP, ‘a’ and ‘b’ represents the lattice parameters. (c) Structural arrangement of the atoms [14].
Figure 1. Characterizations of BP. (a) Multilayer BP. (b) Monolayer BP, ‘a’ and ‘b’ represents the lattice parameters. (c) Structural arrangement of the atoms [14].
Coatings 14 01141 g001
Figure 2. (a) The variation of the bandgap in relation to the thickness of the few-layer phosphorene. (b) The band positions for BP with different layers [18].
Figure 2. (a) The variation of the bandgap in relation to the thickness of the few-layer phosphorene. (b) The band positions for BP with different layers [18].
Coatings 14 01141 g002
Figure 3. The AFM images of BP when exposed to air. Green arrow shows the same region on the flake, and the scale bars are 1 μm [22].
Figure 3. The AFM images of BP when exposed to air. Green arrow shows the same region on the flake, and the scale bars are 1 μm [22].
Coatings 14 01141 g003
Figure 4. XPS spectra of BP after peeling and exposure to air for different time periods. (a) XPS for BP at 0 h, 13 h, 1 day, 2 days, and 3 days ambient exposure. POx peaks appear after 1 day. (b) XPS data for BP on hydrophilic SiO2 and hydrophobic self-assembled OTS on SiO2. Severe BP degradation and POx peaks appear after 24 h for OTS/SiO2, with the BP on SiO2 control sustaining minor degradation in the 24 h. (c) Normalized IR absorbance spectra for BP on OTS/SiO2 versus ambient exposuretime. The P−O stretching mode evolves with lower ambient exposuretime for BP/OTS/SiO2 [22].
Figure 4. XPS spectra of BP after peeling and exposure to air for different time periods. (a) XPS for BP at 0 h, 13 h, 1 day, 2 days, and 3 days ambient exposure. POx peaks appear after 1 day. (b) XPS data for BP on hydrophilic SiO2 and hydrophobic self-assembled OTS on SiO2. Severe BP degradation and POx peaks appear after 24 h for OTS/SiO2, with the BP on SiO2 control sustaining minor degradation in the 24 h. (c) Normalized IR absorbance spectra for BP on OTS/SiO2 versus ambient exposuretime. The P−O stretching mode evolves with lower ambient exposuretime for BP/OTS/SiO2 [22].
Coatings 14 01141 g004
Figure 5. Hyperspectral TEM-EELS spectroscopy of multilayer 2D-Phosphorus Compounds in Air. (a) High-angleannular dark-field (HAADF) contrast image taken at 80 kV; (b) EELS images extracted from the data cube at the energy of phosphorus L2,3-edge (130 eV); (c) EELS spectra corresponding to the selected areas [25].
Figure 5. Hyperspectral TEM-EELS spectroscopy of multilayer 2D-Phosphorus Compounds in Air. (a) High-angleannular dark-field (HAADF) contrast image taken at 80 kV; (b) EELS images extracted from the data cube at the energy of phosphorus L2,3-edge (130 eV); (c) EELS spectra corresponding to the selected areas [25].
Coatings 14 01141 g005
Figure 6. Spectra associated with bandgap regulation of BP. (a) Photo-absorption of bulk phosphorus measured at different optical densities, (black = high; gray = low). (b) Absorbance of 2D phosphorus suspensions that were prepared by fractionation at rcf values near 3000, 5900, 9700, 14,500, and 20,200 g (red to blue). (c) Band structure of bulk black phosphorus at the Z-point of the first Brillouin zone [26].
Figure 6. Spectra associated with bandgap regulation of BP. (a) Photo-absorption of bulk phosphorus measured at different optical densities, (black = high; gray = low). (b) Absorbance of 2D phosphorus suspensions that were prepared by fractionation at rcf values near 3000, 5900, 9700, 14,500, and 20,200 g (red to blue). (c) Band structure of bulk black phosphorus at the Z-point of the first Brillouin zone [26].
Coatings 14 01141 g006
Figure 7. Preparation of BP-based nanocomposites. (a) The preparation process of CuS/BP [37]. (b) the preparation process of BPN/MnO2/DOX [38].
Figure 7. Preparation of BP-based nanocomposites. (a) The preparation process of CuS/BP [37]. (b) the preparation process of BPN/MnO2/DOX [38].
Coatings 14 01141 g007
Figure 8. (a) Synthesis of BP nanosheets by solvothermal method at 60–140 °C [57]. (b) Comparison of the energy band positions of BP nanosheets and the rate of hydrogen production after loading [57].
Figure 8. (a) Synthesis of BP nanosheets by solvothermal method at 60–140 °C [57]. (b) Comparison of the energy band positions of BP nanosheets and the rate of hydrogen production after loading [57].
Coatings 14 01141 g008
Figure 9. (a) BP used as a co-catalyst for photocatalytic H2 production in CdS/BP nanosheets system [59]. (b) Energy diagram and schematic illustration of electron injection from BP NS to TMC [65].
Figure 9. (a) BP used as a co-catalyst for photocatalytic H2 production in CdS/BP nanosheets system [59]. (b) Energy diagram and schematic illustration of electron injection from BP NS to TMC [65].
Coatings 14 01141 g009
Figure 10. (a) Hydrogen production mechanism diagram of BP/Bi2WO6 [69]. (b) Z-scheme photocatalytic water splitting using BP/BiVO4 under visible light irradiation [70].
Figure 10. (a) Hydrogen production mechanism diagram of BP/Bi2WO6 [69]. (b) Z-scheme photocatalytic water splitting using BP/BiVO4 under visible light irradiation [70].
Coatings 14 01141 g010
Table 1. Comparison of the performance of different BP catalysts for photocatalytic hydrogen production.
Table 1. Comparison of the performance of different BP catalysts for photocatalytic hydrogen production.
ApplicationsCatalystsLight SourceScavengerPerformanceReference
Hydrogen evolutionBP>420 nmNa2S/Na2SO366 μmol·h−1·g−1[78]
BP/Pt>420 nm(pH = 6.8)450 μmol·h−1·g−1[57]
BP NSs/Pt>450 nmTriethanolamine89.1 mmol (6 h)[58]
BPNS/Pt(3%)/TMC420/780 nmMethanol/water1.9/0.41 μmol·h−1[65]
BP/MBWO>420 nmTriethanolamine21,042 μmol·g−1 (5 h)[69]
BP/BiVO4>420 nm/160 μmol·h−1·g−1[70]
BP/CdS>420 nmEthanol11,192 μmol·h−1·g−1[71]
0D/2D ZCS/FLP>420 nm/9326 μmol·h−1·g−1[72]
BP/MoS2>420 nmNa2S/Na2SO31286 μmol·h−1·g−1[73]
2D BP/WS2>780 nmEDTA1.55 μmol (3 h)[74]
CdS/BP-MoS2Full spectrumLactic aid183.24 mmol·h−1·g−1[75]
BP/CdS/LTO>420 nmNa2S/Na2SO30.96 mmol·h−1·g−1[76]
BP/Pt/ZnIn2S4>420 nmNa2S/Na2SO31278 μmol·h−1·g−1[77]
BP/Pt/RGO420/780 nmEDTA5.13/1.26 μmol (4 h)[78]
BP/Au/LTO420/780 nmMethanol0.74/0.30 mmol·h−1·g−1[79]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, W.; Yao, B.; Yang, H.; Li, X.; Qiu, L.; Li, S. Application of Metals and Their Compounds/Black Phosphorus-Based Nanomaterials in the Direction of Photocatalytic Hydrogen Evolution. Coatings 2024, 14, 1141. https://doi.org/10.3390/coatings14091141

AMA Style

Zhang W, Yao B, Yang H, Li X, Qiu L, Li S. Application of Metals and Their Compounds/Black Phosphorus-Based Nanomaterials in the Direction of Photocatalytic Hydrogen Evolution. Coatings. 2024; 14(9):1141. https://doi.org/10.3390/coatings14091141

Chicago/Turabian Style

Zhang, Weiwei, Bin Yao, Haotian Yang, Xueru Li, Lina Qiu, and Shaoping Li. 2024. "Application of Metals and Their Compounds/Black Phosphorus-Based Nanomaterials in the Direction of Photocatalytic Hydrogen Evolution" Coatings 14, no. 9: 1141. https://doi.org/10.3390/coatings14091141

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

Article Metrics

Back to TopTop