Next Article in Journal
Catalytic Steam-Reforming of Glycerol over LDHs-Derived Ni–Al Nanosheet Array Catalysts for Stable Hydrogen Production
Next Article in Special Issue
Tuning the Magnetic and Catalytic Properties of Manganese Ferrite through Zn2+ Doping: Gas Phase Oxidation of Octanol
Previous Article in Journal
Electrocatalytic Hydrodechlorination Using Supported Atomically Precise Gold Nanoclusters under Flow-Through Configuration
Previous Article in Special Issue
Influence of Nanoscale Intimacy in Bi-Functional Catalysts for CO2-Assisted Dehydrogenation of C5-Paraffins
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Transition Metal Phosphide Nanocatalysts for H2 Evolution and CO2 Reduction

Nanochemistry Laboratory, Department of Chemistry, Jamia Millia Islamia, New Delhi 110025, India
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(7), 1046; https://doi.org/10.3390/catal13071046
Submission received: 2 June 2023 / Revised: 16 June 2023 / Accepted: 19 June 2023 / Published: 28 June 2023
(This article belongs to the Special Issue Design and Synthesis of Nanostructured Catalysts)

Abstract

:
Green hydrogen energy has captivated researchers and is regarded as a feasible option for future energy-related aspirations. The emerging awareness of renewable energy-driven hydrogen generation and carbon dioxide reduction calls for the use of unconventional schematic tools in the fabrication of nanocatalyst systems. Transition metal phosphides are state-of-art, cost-effective, noble-metal-free materials that have been comprehensively examined for sustainable energy-driven applications. Recent reports on these advanced functional materials have cemented their candidature as high-performance catalytic systems for hydrogen production and for carbon dioxide conversion into value-added chemical feedstock. Bimetallic NiCoP (238.2 mmol g−1 h−1) exhibits top-notch catalytic competence toward photocatalytic HER that reveals the energy-driven application of a pristine class of TMPs, whereas heterostructured Ni2P/CdS was found to be fit for photochemical CO2 reduction, as well as for HER. On the other hand, pristine Ni2P was recently ascertained as an efficient electrocatalytic system for HER and CO2RR applications. A wide array of physicochemical modulations, such as compositional and structural engineering, defect generation, and facet control, have been used for improving the catalytic efficiency of transition metal phosphide nanostructures. In this review, we succinctly discuss the proficiency of transition metal phosphides in green hydrogen production and carbon dioxide conversion via photochemical and electrochemical pathways. We detail the significance of their structural properties and brief the readers about the synthetic advancements without deviating from our goal of summarizing the recent achievements in energy-driven applications.

1. Introduction

The devastating effects of global warming and climate change have motivated researchers to focus their attention beyond fossil fuels and their various modified derivatives. The unchecked release of carbon dioxide (CO2) has drastically perturbed the flora and fauna of Earth’s ecosystems and has put several species on the verge of extinction, as well as threatened the adaptation of many others [1]. Annually, 32.8 billion tonnes of CO2 are released into the atmosphere from the consumption of fossil fuel derivatives such as coal and petroleum [2]. According to the special report proposed by the International Energy Agency (IEA), it is anticipated that the global energy demand will rise by about 25% in the coming two decades because of population expansion, and the energy associated with CO2 release will increase by about 35.8 megatons per year in 2040 [3]. These catastrophic predictions and current environmental maladies necessitate a paradigm shift to a sustainable energy solution. Various accords such as the Kyoto Protocol and the Paris Agreement have been signed to combat CO2 release collectively at multiple levels [4]. The inherent heating nature of CO2 has the potential to cause fluctuations in the seasonal patterns of El Nino and La Nina, which govern the nature of winds worldwide [5]. If standard protocols are not implemented to control the unrestricted venting of CO2, the atmosphere of our planet will be like that of Venus, and the planet will become a non-colonized gas chamber [6]. Consequently, the average global temperature and sea level will increase.
The development of any nation depends heavily on the manufacturing, automotive, and agriculture sectors, but all of these industries require efficient fuels for consequential output. Hence, the optimization of energy resources is economically fundamental for any nation to achieve an equilibrium between the demand and supply chains. Hydrogen (H2), as a renewable energy resource, has various benefits such as remarkable gravimetric density and higher calorific value, which make it a viable alternative to fossil derivatives [7,8,9,10]. Currently, steam reforming of fossil fuels is the standard practice for the generation of bulk H2. However, the nature of this type of H2 is not at all eco-friendly and augments greenhouse gas emission. Therefore, green H2 energy via overall water splitting is sought out as an environmentally benign way to extract H2 [11]. Currently, the research community is committed to elevating the efficiency of the H2 evolution reaction (HER) and CO2 reduction reaction (CO2RR) to the extent that they become a tandem antidote to the climatic concerns of the planet.
The development of HER and CO2RR is hindered by the non-achievement of cost-effectiveness and efficiency in meeting large-scale operational endeavors. HER and the conversion of CO2 into value-added chemicals such as hydrocarbon fuels via photochemical, electrochemical, and photo-electrochemical pathways offer a great deal of cost efficiency and production output because these green routes have been significantly scrutinized [12,13,14,15,16]. Nanocatalysis has proven to be an effective pathway to an environmentally benevolent protocol for CO2 reduction and H2 generation [17]. Nanostructures exhibit significantly higher exposed active sites, and because of quantum confinement effects, they possess optoelectronic properties that are superior to their bulk counterparts [18]. Therefore, exploiting nanocatalysts to carry out HER and CO2RR is instrumental to a scalable production of H2 and carbon-based value-added chemicals. There are numerous fabrication techniques that have been explored for designing the nanostructures of advanced functional materials, such as the reverse micellar system [19,20,21,22], the polymeric precursor route [23], the hydrothermal/solvothermal method [24], etc.
CO2 is an ideal starting material for the production of energetic chemical fuels that are desired to meet energy requirements in the future [25]. Photochemical and electrochemical CO2RR pathways are considered to be economical and environmentally benign routes for the conversion of CO2 into valuable feedstocks such as CO, HCOOH, CH4, C2H4, CH3COCH3, and CH3OH, without producing any undesirable by-products [26,27,28]. The schematic representation of the photocatalytic conversion of CO2 into valuable feedstock is demonstrated in Figure 1. In addition, there is a process of syngas (CO+H2) generation via the photocatalytic/electrocatalytic reduction of CO2 in an aqueous solution that is a vital intermediate for enhancing the yield of hydrocarbon fuels. The reverse water–gas shift reaction (RWGS) is a sustainable pathway of CO2 conversion that has received recognition for its widespread application of CO2 as the starting material [29]. However, all of these HER and CO2RR routes demand Earth-abundant, advanced functional materials that can throttle secondary reactions such as the back recombination of charge carriers, the methanation reaction, or competing HER reactions in the case of CO2 sequestration.
Various classes of materials have been explored as electrocatalysts and photocatalysts, including metals [30], metal alloys [31], metal oxides [32,33], metal sulfides [34], metalorganic frameworks (MOFs), transition metal chalcogenides [35,36], and carbon-based materials like graphene or carbon nanotubes [37,38]. However, metal oxides exhibit low conductivity, and metal sulfides are limited by their photo-corrosiveness. Transition metal phosphides (TMPs) have emerged as efficient catalysts with high electrical conductivity and exceptional physicochemical properties. The catalytic efficiency of TMPs is highly dominated by their intrinsic properties such as light absorptivity, the presence of active sites for the adsorption of reactive species, transportation capability, tunable band potential, and photosensitization. It is relatively difficult to obtain scalable energy-related performance via pristine TMPs, so in order to achieve higher performance, various strategies have been formulated, such as synthetic modulations, cocatalyst incorporation, heterojunction formation, and compositional and morphological modifications. The fabrication of TMPs fundamentally hinges upon the metal-to-phosphorus ratio and the sources of phosphorus. Therefore, researchers are trying to optimize the synthetic methodologies of TMPs that elute non-toxic or less harmful by-products and that are environment friendly in nature. Although TMPs exhibit higher catalytic response for HER and CO2RR applications, there are still several bottlenecks that need to be rectified prior to their large-scale commercial application.
TMPs are an emerging class of advanced materials believed to be potent candidates that can replace precious noble metal-based catalysts [39]. The role of TMPs in nanocatalysis is not only limited to direct utilization. Instead, they can also be employed as cocatalysts. In the last few decades, it has been observed that cocatalyst loading is one of the best ways to enhance photochemical and electrochemical efficiencies by improving their light-harvesting, electrical conductivity, and physicochemical properties through interface engineering, alongside providing multifunctional active sites for surface adsorption [40,41]. Both the activity and selectivity of HER and CO2RR can be enhanced by using cocatalysts. Among different types of cocatalysts, noble metals such as Pt, Pd, Au, Ag, and Ru have been comprehensively studied for the photocatalysis of HER and CO2RR [42,43,44,45]. However, due to their scarcity and the whopping cost involved, their large-scale applicability is limited. An advanced, economical, and efficient method is the need of the hour. Therefore, developing non-noble-metal-based cocatalysts which fulfil the requirement of being cost-effective is requisite. Among various non-noble-metal-based cocatalysts, TMPs are classified as promising candidates for photocatalysis and electrocatalysis. Antil et al. [44] fabricated novel NiCoP@ZnCo-MOF photocatalysts and displayed an augmented rate of H2 (8583.4 μmol g−1 h−1). The group conducted comparative analysis between as-synthesized TMP nanocatalysts relative to the noble-metal-based catalytic system of Pt@ZnCo-MOF (8885.7 μmol g−1 h−1). The rate of H2 production of the TMP-based photocatalyst was found to be nearly analogous to its noble metal counterpart. Thus, we can corroborate TMPs as a viable substitute for noble metals in assisted catalysis for advanced HER applications. An illustrated diagrammatical comparison is provided in Figure 2 for the better selectivity of TMPs over noble metal catalysts.
The schematic landscape of TMPs synthesis, properties, and energy-driven applications such as energy conversion and energy storage are illustrated in Figure 3. In 2005, for the very first time, Liu et al. [46] predicted the catalytic excellence of Ni2P by considering density functional theory and evaluating its photocatalytic activity towards HER. The scope of the CoP cocatalyst has been analyzed to augment the photocatalytic efficiency of two-dimensional g-C3N4. In the CoP/g-C3N4 heterostructured catalyst, photo-generated electron transference occurs from g-C3N4 to the surface of CoP active sites in light [46,47]. The surface electrons react with the absorbed protons, resulting in the supplication of photocatalytic HER activity. Recent reports have also proven the potential of other nickel-based TMPs such as Ni2P, Ni12P5, and Ni3P for enhancing the light absorption capacity of g-C3N4, due to which they are believed to possess high photocatalytic activity in HER [48]. TMPs have achieved superior photocatalytic HER and CO2RR activities, as revealed by the literature review [49]. Hence, researchers need to focus on the developments in the field of TMP-assisted nanocatalysis. This review summarizes the recent advances in TMPs towards HER and CO2RR applications.

2. TMPs Structure and Its Significance

TMPs are versatile catalysts that consist of transition metals bonded with phosphorus (P) atoms. The structure of TMPs is of great significance and can vary depending on the types of metals, the number of P atoms, and the bonding interactions between them. TMPs possess different structures and morphology depending on the other metals and their interactions with the P atoms [50]. In addition, factors like the reactant precursors of transition metal and P, stoichiometry, and synthesis conditions also govern the structural properties of TMPs [51]. In TMPs, the P atom enters the lattice structure of transition metals to form interstitial compounds that act as efficient catalysts for energy conversion processes [52]. The structural modifications on introducing the P atom in the lattice structure of transition metal (M) is an elongation in the bond length of the M-P bond, which results in decreased interactions of M-M bonds and the compression of d-bands [53]. This lattice mismatch tunes the Fermi levels of M, favoring their electronic conductivity and mass diffusion characteristics [54]. TMPs exhibit unique crystal structure where the M atoms form the trigonal prismatic structure, in which the interstitial voids are occupied by P atoms in the smallest structural unit. In contrast, the excess metal ions form nine-fold tetrakaidecahedral structures [55], as represented in Figure 4a,b. The P atom can interact with the transition metal lattice. These structural units can cooperate and have the ability to form miscellaneous lattice types. Variations in the M/P ratio significantly affect the lattice structure of TMPs. Thus, this ratio also controls the catalytic efficiency of TMP nanocatalysts. In a TMP structure, P exhibits relatively higher electronegativity, attributed to the fact it accepts protons readily. This structural feature of TMPs boosts their catalytic action towards HER and CO2RR as facilitated H+/CO2 surface adsorption takes place quickly due to higher electronegativity of P atoms [50]. Therefore, the P content in TMPs is central to determining their catalytic activity. Based on that, they can be classified as transition-metal-rich TMPs (M-rich TMPs) or P-atom-rich TMPs (P-rich TMPs), depending on the M/P ratio [56]. In M-rich TMPs, the M/P ratio is comparatively higher, due to which M–M interactions profoundly affect the physiochemical properties of TMPs. These compounds often exhibit a lower P content than stoichiometric or P-rich TMPs [57]. The excess of M atoms results in unique properties and structures of TMPs. M-rich TMPs are commonly synthesized under high-temperature conditions, such as through the reaction of metal precursors with P sources. However, P-rich TMPs hold a relatively higher ratio of P to transition metal. These compounds have excess P and exhibit unique properties based on their high P content. The crystal structure of P-rich TMPs is depicted in Figure 5a–f below [56]. M-rich TMPs exhibit metallic properties, whereas P-rich TMPs show semiconducting behavior. The metallic behavior in M-rich TMPs is familiar with that of noble metals, showing excellent conducting properties [42]. P-rich TMPs are often synthesized through reactions that involve the addition of excess P or P-rich precursors. For instance, M-rich TMPs bestow higher conductivity and stability than P-rich TMPs due to the larger M–M and M–P bonds than non-conductive P–P bonds.

3. Structural Significance

The structural ascendency of TMPs, due to the introduction of P-atoms in the lattices of transition metal ions, dramatically affects the optoelectronic properties of these interstitial materials, thus making them excellent HER and CO2RR catalysts [58,59,60]. Incorporated P atoms moderately tune the interspace between the transition metal atoms that restricts their molecular interaction. It leads to the shrinking of d bands and increasing the concentration of energy levels in the vicinity of the Fermi center [61]. This excellent optoelectronic modulation in TMPs escalates their catalytic efficiency of the order of noble-metal based catalytic systems [62,63]. The structural engineering of TMPs is one of the finest strategic tools to ameliorate their performance towards HER and CO2RR applications. Recent reports of TMPs support the fact that developing heterojunctions, hollow and nanoarray structures, significantly enhance their scope of applicability. Heterojunction formation allows for the interfacial charge transfer during photochemical HER and CO2RR, which effectively subdues the back recombination reactions and assures the utilization of active electron carriers in primary reduction reactions. On the other hand, the formation of hollow structures eases the adsorption pathways of reactant species by providing higher exposed active sites available for surface reactions and sufficient mass diffusion routes [64,65,66,67,68]. Higher surface areas and surface roughnesses in nanoarrays facilitate the contact compass between catalytic system and adsorbed species. The structural properties of TMPs are pretty significant in modulating their electronic properties, leading to remarkable catalytic responses in energy-related applications. Other than structural engineering, tuning the physicochemical properties of TMPs also hinges on their compositional and morphological properties.

4. Advances in TMP Fabrication

For the comprehensive exploitation of TMPs, optimizing synthetic routes is essential to achieve cost-effectiveness and environmental friendliness. Conventional methods used to fabricate TMPs, such as temperature-programed reduction under H2 atmosphere or the ball milling method, are restricted to the synthesis of bulk TMPs only. However, for constructing nano-dimensional TMPs, fabrication routes can be classified on the basis of P sources used and the different methods involved. Synthetic pathways have been classified based on P sources such as organic phosphines (trioctylphosphine), hypophosphites (NH4H2PO2, NaH2PO2), elemental phosphorus (white, red), and phosphorenes [13]. Phosphorenes are the exfoliated monolayers of black P that have also been utilized as P precursors for TMP engineering [69]. One different approach to designing TMPs is the pyrolysis of P and carbon-based precursors alongside the source of the transition metal ion. Integrating carbon-based materials during TMP synthesis impedes the agglomeration of TMP nanoparticles and accelerates their conductivity. Therefore, different morphologies and particle sizes can be achieved based on the use of varying P sources for TMP nanostructures. The discrete synthesis method for designing TMPs can vary metal to metal or depending on the desired morphology, the targeted energy-related application, and the chosen precursors of P [70]. Researchers often optimize the reaction conditions and other parameters such as reducing the agent and solvent system to obtain the desired physicochemical properties and morphology of the prepared TMP nanocatalysts [71]. Synthesizing TMP nanostructures typically involves diverse physical and chemical routes such as solid-state reactions [72], the phosphorization of metal precursors [73,74,75,76], chemical vapor deposition (CVD) [77], and hydrothermal pathways [78,79,80,81,82].
In the usual practice, the solid-state method proceeds via heating a mixture of transition metal precursors (metallic powder or metal oxides) alongside a P source at significantly higher temperatures under an inertial atmosphere to subdue any undesired defects or oxide impurities [72]. The reaction advances through a series of intermediate steps that lead to the development of the desired TMP nanostructures. As its name suggests, the phosphorization of metal precursors comprises the precursor of the desired metal and is minted with the source of P, such as red or white phosphorus. The phosphorization reaction can occur either in solution or in the gas phase, depending on the nature of transition metal precursor and P source. The obtained reaction mixture is then further processed to isolate the pristine TMPs without any foreign impurities. The CVD technique is generally utilized to deposit thin films or coatings of TMPs. In this pathway, relatively volatile metallic precursor is integrated with the P-containing precursor, and the resulting mixture is introduced into the reaction chamber, followed by that reactant mixture decomposed at high temperatures leading to the deposition of the TMPs onto a substrate. The hydrothermal/solvothermal route is an environmentally benign chemical route that offers great deal of output and control over morphology for far-reaching applications of TMP nanostructures [80]. It involves the reaction of metal precursors with the P source in generally aqueous media and sometimes in other solvents under high-pressure and high-temperature conditions. The organic solvent environment is developed in the reaction media to achieve the desired morphology. The reaction is typically carried out in an autoclave, where the reaction mixture is subjected to the particular reaction conditions. The main advantage of procuring this route is the formation of well-defined TMP nanoparticles. Considering the aforementioned synthetic routes of TMPs, we manifest that all these pathways have their respective advantages and disadvantages. However, in our vision, the hydrothermal route has far-reaching applications for the sake of energy-related operations of TMP-based hybrid materials.

5. TMP Photocatalysts for HER and CO2RR Applications

Over the past few years, the optoelectronic properties and chemical stability of TMPs have made them promising catalytic systems for photochemical HER and CO2RR applications. TMPs such as FeP [83], CoP [84], NiCoP [85], MoP [86], and Cu3P [87] have been investigated in the field of nanocatalysis and are known to display exceptional activity and selectivity towards HER and CO2RR applications. There are various methods of enhancing the photocatalytic performances of TMPs such as metallic/non-metallic doping [88], surface modulation [89], and mixing with other nanomaterials to form heterojunctions [90]. For instance, the integration of noble metals like Au and Pt with TMPs ameliorates their photocatalytic activity towards HER and CO2RR [49,91]. In the past few decades, much of the attention has been placed on improving the efficiency of photocatalytic processes by developing and probing miscellaneous catalytic systems. Therefore, to achieve this goal, researchers are persistently trying to enhance the activity and selectivity of TMP-based photocatalysts. However, this transformation is limited by the quick recombination rate of photogenerated electron–hole pairs and the restricted photo-sensitization of light absorbers during photocatalytic operations. These two obstacles are the major bottlenecks of photocatalytic HER and CO2RR operations that impede their quantum efficiency in achieving H2 and carbonaceous value-added chemicals [49]. The tailored band energy of TMP nanocatalysts allows an accelerated harvesting of light that ultimately leads to the flow of uninterrupted photo-induced charge carriers, allowing longer runs of photocatalysis. The prevention of the back recombination of photo-generated electron–hole pairs can be successfully achieved by developing appropriate cocatalysts that would generate suitable interfaces, such as Schottky junctions, S-schemes, Z-schemes, p-n heterojunctions, and type II heterojunctions, leading to the fast movement of charge carriers from one component to the other, thus reducing the probability of electron–hole pair recombination [92,93,94]. Although, some secondary side reactions result in the photocatalytic CO2RR. HER is one of the competent reactions in the conversion of CO2 to CO. Therefore, the recent trends of TMP-assisted photocatalytic applications have imbibed these limiting factors, as demonstrated by the reports of TMPs discussed hereafter in this section.
Li et al. [78] fabricated a Ni2P/ZnIn2S4 heterostructure to examine its scope towards photocatalytic water splitting. Optimized Ni2P/ZnIn2S4 heterojunctions exhibited a remarkable H2 evolution rate of ~2.1 mmol g−1 h−1, which can be attributed to the better synergism between the photocatalytic components, the enhanced charge separability, the larger surface area, and the outstanding electron transport ability. Shen et al. [95] synthesized EosinY-Cu3P-CNT (carbon nanotubes). The group reported an exceptional H2 evolution rate of 17.22 mmol g−1 h−1, which is owed to the efficient separation of charge carriers in Eosin Y and enhanced electron transfer ability that symbiotically boosted the rate of H2 generation in the TMP-based photocatalyst. Meng et al. [40] fabricated Ru-CoP-1:8/GCN (g-C3N4) photocatalysts via the wet chemical reduction method for efficient H2 production. The reported study showed an exceptional rate of 1172.5 μmol g−1 h−1. This report outlines the cost-effectiveness of Ru-CoP heterostructured catalytic system as compared to noble metal catalysts for photocatalytic applications. Gong et al. [86] developed Ni-MoP@NPPF (porous nitrogen-doped carbon nanofibers) via electrospinning followed by phosphorization as well as carbonization and showed the rate of CO production to be 953.53 μmol g−1 h−1. It exhibited an impressive selectivity for CO of 98.95% due to the formation of Ni-N bonds. The mechanism proposed that the transfer of electrons occurs across the interface between the two catalysts, enhancing the rate of CO2 reduction, as represented in Figure 6a. The photocatalyst revealed exceptional activity and selectivity for CO2 reduction with Ni, as without Ni co-catalyst, the yield was found to be relatively inferior, as depicted in Figure 6b. The product selectivity of Ni-MoP@NCPF is realized by higher CO generation, as the optimized Ni-MoP@NCPF catalytic system exhibited greatly enhanced photocatalytic CO2 reduction performance towards CO generation compared to CH4 and H2 product formation, as depicted in Figure 6c. Ni-MoP@NCPF provided channels for CO2 diffusion and electron transport. This type of TMP-based catalytic system is the epitome of collaborative morphological, compositional, and hetero-interfacial engineering, as the distribution of Ni throughout the MoP@NCPF improved the transformation of CO2 to CO. The main factors that regulate the degree of photocatalytic HER and CO2RR processes are the separation, transfer, and recombination of electron–hole pairs. The photocatalytic activity of CO2 conversion is ameliorated by the higher separation rate of photo-induced charge carriers or lower recombination rate ascribed to more electrons available for primary reduction reactions at the conduction band of TMP-based photocatalysts.
Sun et al. [79] fabricated Cu3P-Ni2P/g-C3N4 nanocatalysts via a solvothermal pathway to determine their photochemical HER efficiency. The H2 production rate was found to be considerably efficient due to the formation of p-n heterojunctions, assisting in the facile charge transfer and exposure of a greater number of active sites. Figure 7a,b contain mechanistic sketches of Cu3P-Ni2P/g-C3N4-assisted photocatalytic H2 generation and the transfer of electrons via heterojunctions after the advent of an internal electric field before and after visible light irradiation, respectively. Song et al. [96] fabricated WP-NC (nitrogen-doped carbon)/g-C3N4 heterojunctions for efficient H2 generation via photocatalytic and electrocatalytic pathways. The optimized WP-NC/g-C3N4 catalyst exhibited an H2 evolution rate as high as 1.2 mmol g−1 h−1, owing to the deposition of NC layers over WP, which ameliorated the charge separation and transfer efficiency during photocatalytic water splitting.
Su et al. [80] engineered Ni2P/CdS photocatalysts via the hydrothermal pathway and studied their efficiency for HER and CO2RR applications. The resultant band gap of the Ni2P/CdS photocatalyst was found to be 2.16 eV with multiple production rate of H2, CO, and CH4 generation with porous and non-porous nanocomposites, as depicted in Figure 8. In this report, heterojunction formation played an essential role in the enhanced photocatalytic efficiency of Ni2P/CdS, ascribed to the facilitated charge transfer. Guo et al. [83] synthesized an FeP/CN photocatalyst via the thermal decomposition method and reported the maximum rate of CO production to be 5.19 μmol g−1 h−1. The synthesized catalysts exhibited the tremendous conversion efficiency of CO2 to CO because of the greater active sites and facile charge transfer during the photocatalytic CO2RR. Wang et al. [97] designed and synthesized novel Fe-doped CoP for CO2 reduction to CO. The results reported a maximum selectivity of 90.3% (CO), and the rate of evolution was estimated to be 21 μmol h−1. Fe doping assisted in subduing the activation energy barrier for intermediate formation and promoted CO evolution. Lv et al. [98] synthesized ultrathin NiCoP nanosheets to achieve an effective rate of HER (238.2 mmol g−1 h−1). The reported bimetallic catalysts exhibited bifunctional properties such as electrocatalysis and photocatalysis efficiently. Duo et al. [99] designed FeP/CdS heterostructured photocatalysts via the solvothermal route, and the as-prepared nanocatalyst produced H2 at the rate of 37.92 mmol g−1 h−1. This excellent HER was attributed to the charge separation and better electron transportation demonstrated by the heterojunction formation between the FeP and CdS. Li et al. [72] fabricated Ni2P/NiO/CN (graphitic carbon nitride) via solid–gas reaction for the conversion of CO2 to CO and CH4. The enhanced activity resulted from the formation of a p-n heterojunction between NiO and CN, while Ni2P governed the activation and adsorption of CO2 reactant species. The catalytic proficiency of TMPs towards photochemical HER and CO2RR applications are tabulated in Table 1.

6. TMP Electrocatalysts for HER via Water Splitting

Over the last few years, a colossal number of attempts have been dedicated to the advancement of TMP-based electrocatalysts for propelling the applications of HER, and considerably accomplishments have been achieved by ameliorating the final results through heteroatom doping and nanostructure engineering [100,101,102]. Li et al. [74] synthesized Co2P/Ni2P nanohybrids for HER via water splitting. The as-developed nanohybrids required a very low overpotential (51 mV) to achieve 10 mA cm−2 alongside noteworthy operational durabilities. This catalytic performance exceeded most of the reported non-noble TMP-based electrocatalysts. The exceptional results could be attributed to the large surface area, an abundance of active sites, and robust interfacial contact between Ni2P and Co2P. In another reported study, Cho et al. [75] examined the HER activity of Co-, Ni-, and Mn-doped FeP nanoparticles. The electrochemical results showed that Co-FeP nanoparticles required a 126 mV overpotential to attain 10 mA cm−2 and exhibited higher cathodic current density than Ni- and Mn-doped FeP because of their fast charge transfer rate and high, electrochemically active surface area. In another study, Co-, Fe-, Mn-, Na-, Cr-, Li-, V-, Nb-, Ti-, Pb-, and Sn-doped Ni2P electrocatalysts were investigated by Xiong et al. [76] for their HER catalytic activity. The electrochemical results demonstrated that Fe- and Co-doped Ni2P exhibited superior performances similar to Pt-like activity with a very low overpotential of 31 mV at 10 mA cm−2. These results are ascribed to the charge redistribution on the catalyst’s surface, produced by the doping effect. Li et al. [103] fabricated MoP/MoNiP@C heterostructures to examine their electrochemical HER activity, and as a result, the as-prepared electrocatalyst showed remarkable performance, with a 134 mV overpotential and a 66 mV dec−1 Tafel slope. These excellent results can be ascribed to the synergistic effect between MoNiP and MoP nanoparticles that augmented the active sites of the catalyst. The LSV (linear sweep voltammetric) curves and Tafel plots of as-prepared electrocatalysts towards HER are illustrated in Figure 9a,b, whereas cyclic voltammetric (CV) curves and a double-layer capacitance plot of MoP/MoNiP@C are demonstrated in Figure 9c,d. As shown in Figure 9e, the MoP/MoNiP@C impedance is relatively smaller than other electrocatalysts, which corroborates the higher catalytic performance and outstanding conductivity of MoP/MoNiP@C. To examine the robustness of MoP/MoNiP@C for longer runs, stability cycles of the LSV curves were recorded for 2000 cycles, and minor changes in the cathodic curve were observed, as shown in Figure 9f. The onset of Figure 9f depicts the time-dependent current density curve of the optimized electrocatalyst. The fabrication flowchart of MoP/MoNiP@C is illustrated in Figure 10.
Yang et al. [104] designed Ni2P/MoO2/NF nanorods-type heterostructured electrocatalyst for HER water splitting. As-prepared electrocatalysts exhibited very low overpotential (34 mV), which was attributed to the combined effect between MoO2 and Ni2P that was enhanced due to the robust electronic coupling effect. In a reported work investigating the HER activity, Kang et al. [105] synthesized NiFeP@C electrocatalyst that showed notable results, including the low overpotential of 160 mV towards HER. The enhanced results accredited to the synergistic effect between P, Fe, Ni, and C, which accelerated the pace of charge transfer and escalated the electrocatalytic performance. Chen et al. [106] fabricated Ru-MnFeP/NF electrocatalysts to examine HER performance. The results showed a low overpotential of 35 mV alongside high stability for 50 h. Liu et al. [107] engineered Mn- and Ni-deposited FeP (Ni-Mn-FeP) electrocatalyst and investigated its HER activity. The as-prepared electrocatalyst demonstrated enhanced results with 103 mV ultralow overpotential. The combination technique of co-doped high-valence and low-valence metals motivated the advancement of high-activity and functional catalysts. In a reported study, Er-doped NiCoP/NF nanowires were fabricated by Zhang et al. [108] to assess their HER activity. The electrocatalytic efficiency of the as-prepared electrocatalyst was ascribed to its lowered overpotential value of 46 mV to reach 10 mA cm−2 cathodic current density. The advanced performance of Er-doped NiCoP/NF was ascribed to the amalgamation of Er to NiCoP that considerably adjusted d-band centers alongside electronic structure of Co and Ni atoms by altering to lower energies with regards to Fermi level and also enhanced the HER/OER intermediate Gibbs free energies.

7. TMP Electrocatalysts for CO2RR Application

TMPs have emerged as multi-facet materials with fascinating electronic and structural features, which have led to impeccable catalytic activity for critical energy transformation such as CO2 reduction [109,110]. There are several recent reports of electrochemical CO2RR applications of TMP-based nanocatalysts. Downes et al. [111] synthesized Cu3P nanoparticles to investigate their electrochemical CO2RR performance. The results showed the formation of formate with as high as 8% Faradaic efficiency (FE). The enhanced activity and stability were accredited to using a solution-based molecular precursor method for Cu3P fabrication that offered multi-dimensional opportunities for altering the morphology, surface chemistry, and composition to attain notable CO2 conversion efficiency. In another work, Ji et al. [82] incorporated FeP on a Ti mesh (FeP/TM) that functioned as an effective electrocatalyst for CO2 reduction to transform into alcohols with 94.3% FECH3OH+C2H5OH and 80.2% FECH3OH. The improved results were attributed to the combining effect between two adjacent Fe atoms, as revealed by density functional theory (DFT) calculations. Sun et al. [112] designed MoP/In-PC for electrocatalytic CO2 reduction. The current density and FE of the as-designed electrocatalyst reached 43.8 mA cm−2 and 96.5%, respectively. These noteworthy results were accredited to the advantageous feature of nanosized MoP, the adsorption ability of the robust CO2 and CO2-intermediate, the high interfacial charge transfer, and the combining effect between In-doped carbon supports and MoP. Kim et al. [113] synthesized Ni2P/Ho2O3 core–shell nanoparticles (CSNPs) to examine their scope in electrochemical CO2RR applications. As-synthesized electrocatalysts generated acetone with 25.4% FE that was attributed to the synergistic effect between Ni2P and Ho2O3. Figure 11 depicts the synthesis steps for designing Ni2P/Ho2O3 catalytic systems. The four main value-added products in the form of H2, HCOOH, CH3OH, and (CH3)2CO resulted from the reduction of CO2, as revealed in the Figure 12a, Ni2P/Ho2O3 core–shell nanoparticles (CSNPs) have been found to be very beneficial for the electrocatalytic behavior towards (CH3)2CO. While Ho2O3 nanodisks (NDs) produced H2, HCOOH, and CH3OH, only H2 was generated by Ni2P in electrocatalytic CO2 reduction towards C2 or C3 pathways, as shown in Figure 12b–d.
Laursen et al. [114] fabricated Cu3P NS/Cu for electrocatalytic CO2RR. The results showed the formation of formate with a 65 mV lower overpotential value and 0.9% FE. A Fe2P electrocatalyst was synthesized by Calvinho et al. [115] to examine the CO2 conversion efficiency. The outcomes showed a 53% FE with ethylene glycol (C2) product formation after CO2 reduction, which inferred successful C-C coupling during the surface adsorption reaction mechanism. Banerjee et al. [116] fabricated Ni2P electrocatalyst for the reduction of CO2, and the result demonstrated the formation of formaldehyde along with the surface hydrogen affinity and dynamic reconstruction of the surface through adsorption of H that promoted the C–C coupling and selective reduction of CO2 on Ni2P. The catalytic proficiency of TMPs towards electrochemical HER and CO2RR applications are tabulated in Table 2 and Table 3, respectively.
The aforementioned path-breaking reports are proof for the ongoing research on TMPs as cocatalyst or electrocatalyst in HER via water splitting and CO2RR applications. However, there are still many mysteries unsolved regarding the synthesis of TMPs that comprise tail gas post-treatment and unstable reactants, which create obstacles for the bright future of TMPs as electrocatalysts. Therefore, coping with these bottlenecks of TMP nanostructures can simplify their success in energy-driven catalytic applications at scalable points.

8. Future Perspectives

The outlook of nanocatalysis-assisted HER and CO2RR is propitious thanks to wide array of physicochemical modulations that can be carried out in different nanomaterials. Alleviating the cost of green HER and CO2RR applications is fundamental for applying these sustainable pathways to a scalable utility in day-to-day life. Therefore, designing cheap catalytic systems as an economical substitute for noble metal catalysts is the need of the hour in order to cope up with the latter’s whopping costs, which currently hinders the widespread exploitation of green HER and CO2RR applications [117,118,119,120]. In the last few years, researchers have realized the potential of TMPs as tremendously efficient catalytic systems able to address the energy related applications at scalable points [121,122,123,124]. Extensive energy-focused research operations have been performed by employing TMP nanostructured catalytic systems [125]. TMPs are emerging advanced functional materials that have recently exhibited great potential in catalyzing HER and CO2RR to generate value-added chemicals and fuels [126,127,128]. On account of their versatile physicochemical properties, TMPs are beneficial in a broad range of applications, varying from catalysis to energy storage devices. TMPs come forth as remarkable catalytic systems which offer a high range of performances and great robustness. Their inexpensiveness and recent ground-breaking advances promise to fill the void between research and commercial applicability.
Experimental and computational investigations of TMP nanostructures highlight their marvelous catalytic efficiency, stability, and selectivity towards HER and CO2RR applications [129]. However, the extent of their catalysis hinges upon exposed active sites, which govern the surface adsorption reactions. Therefore, it is fundamentally important to enhance the interaction between TMPs’ surface sites and the reactant molecules to escalate the turnover frequency of generating H2 and another chemical feedstock [130]. One major bottleneck of TMP-assisted nanocatalysis is the scarcity of reports which systematically explore and elucidate the reaction mechanism behind their inherent catalytic proficiency at an atomic scale. To examine the in-situ behavior of TMPs, sophisticated characterization techniques such as X-ray absorption and operando and in situ Raman spectroscopies can be upgraded in accordance with the reaction conditions for probing the intermediate steps involved during HER and CO2RR applications. In addition, theoretical models should be developed to determine the thermodynamic landscape and kinetics of surface reactions on the active TMPs centers.

9. Conclusions

Herein, we have discussed the exigency of renewable energy-driven H2 production and CO2 sequestration in order to meet the current energy demand through clean energy resources. TMP nanocatalysts have been reviewed as heterogeneous catalytic systems for HER and CO2RR applications. The structural advantages of TMPs have been discussed precisely, as structural engineering is one of the most sought-after pathways for enhancing their activity and stability. To understand the structural significance of TMPs, the role of P and M/P ratio was reviewed. We have summarized the most utilized synthetic methodologies for designing TMP nanostructures. Photocatalytic HER and CO2RR applications were surveyed collectively under the enlightenment of recent achievements of TMPs. Bimetallic NiCoP exhibits an enormous rate of H2 production (238.2 mmol g−1 h−1), and heterostructured CoP/rGO was found to have a higher rate of reducing CO2 to CO (47.33 mmol g−1 h−1). Correspondingly, recent advances in electrocatalytic HER and CO2RR applications were also reviewed, which corroborate the diverse applicability of TMP nanocatalysts in energy-driven applications. Pristine Ni2P has emerged as a resourceful electrocatalyst for HER. Nevertheless, there are only a handful reports for electrocatalytic CO2 reduction, which is at the cutting edge of technology and needs to be developed further in future environmental aspirations. For CO2RR applications, TMPs can be exploited in either gas-phase or liquid-phase reactions, which facilitate the conversion of CO2 into chemical feedstocks such as C1 and C2 products and other higher hydrocarbons. The P content and physicochemical properties of TMPs play major roles in regulating their catalytic responses. Therefore, optimizing reaction conditions and synthetic routes is instrumental in developing these emerging functional materials for energy-related applications. TMPs can be a game-changer in the field of nanocatalysis thanks to their Earth-abundance and cost-effectiveness.

Author Contributions

S.S. is responsible for the preparation of the rough manuscript by surveying different reports and literature. S.A.A. collected the data and statistics from recent reports. U.F.M. is responsible for compiling the rough draft into manuscript form. I.S. refined the manuscript. T.A. is responsible for the conceptualization, supervision, accruing funding, analysis, and finalization of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

T.A. thanks UGC, New Delhi, Government of India, for the Research Grant for In-service Faculty Members. The authors also thank SERB, CSIR, and MoE (SPARC/2018-2019/P843/SL) for financial support to Nano Chemistry and Nano Energy Labs.

Data Availability Statement

Not applicable.

Acknowledgments

T.A. thanks the various research schemes sponsored by SERB, BRNS, CSIR, and MoE-SPARC, Government of India, for the financial support to Nano Chemistry and Nano Energy Labs at Jamia Millia Islamia. T.A. also thanks UGC, New Delhi, for the Research Grant for In-service Faculty Members. S.S., S.A.A. and I.S. thank UGC, New Delhi, for the Research Fellowships.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ali, S.A.; Sadiq, I.; Ahmad, T. Oxide based Heterostructured Photocatalysts for CO2 Reduction and Hydrogen Generation. ChemistrySelect 2023, 8, 202203176. [Google Scholar] [CrossRef]
  2. Moutinho, V.; Madaleno, M.; Inglesi-Lotz, R.; Dogan, E. Factors affecting CO2 emissions in top countries on renewable energies: A LMDI decomposition application. Renew. Sustain. Energy Rev. 2018, 90, 605–622. [Google Scholar] [CrossRef] [Green Version]
  3. Waheed, R.; Chang, D.; Sarwar, S.; Chen, W. Forest, agriculture, renewable energy, and CO2 emission. J. Clean. Prod. 2018, 172, 4231–4238. [Google Scholar] [CrossRef]
  4. Leggett, J.A. The United Nations Framework Convention on Climate Change, the Kyoto Protocol, and the Paris Agreement: A Summary; UNFCC: New York, NY, USA, 2020; p. 2. [Google Scholar]
  5. Chiodi, A.M.; Harrison, D.E. Comment on Qian et al. La Niña and El Niño composites of atmospheric CO2 change. Tellus B Chem. Phys. Meteorol. 2014, 66, 20428. [Google Scholar] [CrossRef] [Green Version]
  6. Ali, S.A.; Ahmad, T. Treasure trove for efficient hydrogen evolution through water splitting using diverse perovskite photocatalysts. Mater. Today Chem. 2023, 29, 101387. [Google Scholar] [CrossRef]
  7. Shaheen, S.; Sadiq, I.; Ali, S.A.; Ahmad, T. Bismuth-Based Multi-Component Heterostructured Nanocatalysts for Hydrogen Generation. Catalysts 2023, 13, 295. [Google Scholar] [CrossRef]
  8. Abdalla, A.M.; Hossain, S.; Nisfindy, O.B.; Azad, A.T.; Dawood, M.; Azad, A.K. Hydrogen production, storage, transportation and key challenges with applications: A review. Energy Convers. Manag. 2018, 165, 602–627. [Google Scholar] [CrossRef]
  9. Naaz, F.; Alshehri, S.M.; Mao, Y.; Ahmad, T. Unraveling the chemoselective catalytic, photocatalytic and electrocatalytic applications of copper supported WO3 nanosheets. Catal. Commun. 2023, 178, 106678. [Google Scholar] [CrossRef]
  10. Sadiq, I.; Ali, S.A.; Ahmad, T. Graphene-Based Derivatives Heterostructured Catalytic Systems for Sustainable Hydrogen Energy via Overall Water Splitting. Catalysts 2023, 13, 109. [Google Scholar] [CrossRef]
  11. Singla, S.; Sharma, S.; Basu, S.; Shetti, N.P.; Aminabhavi, T.M. Photocatalytic water splitting hydrogen production via environmental benign carbon-based nanomaterials. Int. J. Hydrogen Energy 2021, 46, 33696–33717. [Google Scholar] [CrossRef]
  12. Ramadoss, M.; Chen, Y.; Chen, X.; Su, Z.; Karpuraranjith, M.; Yang, D.; Pandit, M.A.; Muralidharan, K. Iron-modulated three-dimensional CoNiP vertical nanoarrays: An exploratory binder-free bifunctional electrocatalyst for efficient overall water splitting. J. Phys. Chem. C 2021, 125, 20972–20979. [Google Scholar] [CrossRef]
  13. Shi, Y.; Zhang, B. Recent advances in transition metal phosphide nanomaterials: Synthesis and applications in hydrogen evolution reaction. Chem. Soc. Rev. 2016, 45, 1529–1541. [Google Scholar] [CrossRef] [PubMed]
  14. Dong, Y.; Kong, L.; Jiang, P.; Wang, G.; Zhao, N.; Zhang, H.; Tang, B. A general strategy to fabricate NixP as highly efficient cocatalyst via photoreduction deposition for hydrogen evolution. ACS Sustain. Chem. Eng. 2017, 5, 6845–6853. [Google Scholar] [CrossRef]
  15. Landers, A.T.; Fields, M.; Torelli, D.A.; Xiao, J.; Hellstern, T.R.; Francis, S.A.; Tsai, C.; Kibsgaard, J.; Lewis, N.S.; Chan, K.; et al. The predominance of hydrogen evolution on transition metal sulfides and phosphides under CO2 reduction conditions: An experimental and theoretical study. ACS Energy Lett. 2018, 3, 1450–1457. [Google Scholar] [CrossRef] [Green Version]
  16. Li, Y.; Sun, Y.; Qin, Y.; Zhang, W.; Wang, L.; Luo, M.; Yang, H.; Guo, S. Recent advances on water-splitting electrocatalysis mediated by noble-metal-based nanostructured materials. Adv. Energy Mater. 2020, 10, 1903120. [Google Scholar] [CrossRef]
  17. Stolarczyk, J.K.; Bhattacharyya, S.; Polavarapu, L.; Feldmann, J. Challenges and prospects in solar water splitting and CO2 reduction with inorganic and hybrid nanostructures. ACS Catal. 2018, 8, 3602–3635. [Google Scholar] [CrossRef]
  18. Pandit, N.A.; Ahmad, T. Tin Oxide Based Hybrid Nanostructures for Efficient Gas Sensing. Molecules 2022, 27, 7038. [Google Scholar] [CrossRef]
  19. Vaidya, S.; Ahmad, T.; Agarwal, S.; Ganguli, A.K. Nanocrystalline oxalate/carbonate precursors of Ce and Zr and their decompositions to CeO2 and ZrO2 nanoparticles. J. Am. Ceram. Soc. 2007, 90, 863–869. [Google Scholar] [CrossRef]
  20. Ahmad, T.; Ganguli, A.K. Structural and dielectric characterization of nanocrystalline (Ba, Pb) ZrO3 developed by reverse micellar synthesis. J. Am. Ceram. Soc. 2006, 89, 3140–3146. [Google Scholar] [CrossRef]
  21. Ahmad, T.; Shahazad, M.; Ubaidullah, M.; Ahmed, J.; Khan, A.; El-Toni, A.M. Structural characterization and dielectric properties of ceria–titania nanocomposites in low ceria region. Mater. Res. Express 2017, 4, 125016. [Google Scholar] [CrossRef]
  22. Ganguli, A.K.; Vaidya, S.; Ahmad, T. Synthesis of nanocrystalline materials through reverse micelles: A versatile methodology for synthesis of complex metal oxides. Bull. Mater. Sci. 2008, 31, 415–419. [Google Scholar] [CrossRef]
  23. Ahmad, T.; Lone, I.H.; Ubaidullah, M. Structural characterization and multiferroic properties of hexagonal nano-sized YMnO3 developed by a low temperature precursor route. RSC Adv. 2015, 5, 58065–58071. [Google Scholar] [CrossRef]
  24. Fazil, M.; Ahmad, T. Pristine TiO2 and Sr-Doped TiO2 Nanostructures for Enhanced Photocatalytic and Electrocatalytic Water Splitting Applications. Catalysts 2023, 13, 93. [Google Scholar] [CrossRef]
  25. Zhang, M.; Xu, Y.; Williams, B.L.; Xiao, M.; Wang, S.; Han, D.; Sun, L.; Meng, Y. Catalytic materials for direct synthesis of dimethyl carbonate (DMC) from CO2. J. Clean. Prod. 2021, 279, 123344. [Google Scholar] [CrossRef]
  26. Chauhan, D.K.; Sharma, N.; Kailasam, K. A critical review on emerging photocatalysts for syngas generation via CO2 reduction under aqueous medium: A sustainable paradigm. Adv. Mater. 2022, 3, 5274–5298. [Google Scholar] [CrossRef]
  27. Zaza, L.; Rossi, K.; Buonsanti, R. Well-defined copper-based nanocatalysts for selective electrochemical reduction of CO2 to C2 products. ACS Energy Lett. 2022, 7, 1284–1291. [Google Scholar] [CrossRef]
  28. Kim, D.; Kley, C.S.; Li, Y.; Yang, P. Copper nanoparticle ensembles for selective electroreduction of CO2 to C2–C3 products. Proc. Natl. Acad. Sci. USA 2017, 114, 10560–10565. [Google Scholar] [CrossRef] [Green Version]
  29. Tu, W.; Zhou, Y.; Zou, Z. Photocatalytic conversion of CO2 into renewable hydrocarbon fuels: State-of-the-art accomplishment, challenges, and prospects. Adv. Mater. 2014, 26, 4607–4626. [Google Scholar] [CrossRef]
  30. Khan, H.; Lone, I.H.; Lofland, S.E.; Ramanujachary, K.V.; Ahmad, T. Exploiting multiferroicity of TbFeO3 nanoparticles for hydrogen generation through photo/electro/photoelectro-catalytic water splitting. Int. J. Hydrogen Energy 2023, 48, 5493–5505. [Google Scholar] [CrossRef]
  31. Liu, L.; Corma, A. Structural transformations of solid electrocatalysts and photocatalysts. Nat. Rev. Chem. 2021, 5, 256–276. [Google Scholar] [CrossRef]
  32. Fazil, M.; Alshehri, S.M.; Mao, Y.; Ahmad, T. Hydrothermally Derived Mg-Doped TiO2 Nanostructures for Enhanced H2 Evolution Using Photo-and Electro-Catalytic Water Splitting. Catalysts 2023, 13, 893. [Google Scholar] [CrossRef]
  33. Moon, S.Y.; Gwag, E.H.; Park, J.Y. Hydrogen Generation on Metal/Mesoporous Oxides: The Effects of Hierarchical Structure, Doping, and Co-catalysts. Energy Technol. 2018, 6, 459–469. [Google Scholar] [CrossRef] [Green Version]
  34. Mamiyev, Z.; Balayeva, N.O. Metal Sulfide Photocatalysts for Hydrogen Generation: A Review of Recent Advances. Catalysts 2022, 12, 1316. [Google Scholar] [CrossRef]
  35. Zhang, H.; Li, J.; Tan, Q.; Lu, L.; Wang, Z.; Wu, G. Metal–organic frameworks and their derived materials as electrocatalysts and photocatalysts for CO2 reduction: Progress, challenges, and perspectives. Chem. Eur. J. 2018, 24, 18137–18157. [Google Scholar] [CrossRef]
  36. Ali, S.A.; Ahmad, T. Chemical strategies in molybdenum-based chalcogenides nanostructures for photocatalysis. Int. J. Hydrogen Energy 2022, 47, 29255–29283. [Google Scholar] [CrossRef]
  37. Kuang, P.; Sayed, M.; Fan, J.; Cheng, B.; Yu, J. 3D graphene-based H2-production photocatalyst and electrocatalyst. Adv. Energy Mater. 2020, 10, 1903802. [Google Scholar] [CrossRef]
  38. Mehtab, A.; Alshehri, S.M.; Ahmad, T. Photocatalytic and photoelectrocatalytic water splitting by porous g-C3N4 nanosheets for hydrogen generation. ACS Appl. Nano Mater. 2022, 5, 12656–12665. [Google Scholar] [CrossRef]
  39. Hameed, R.A. Nanostructured Phosphides as Electrocatalysts for Green Energy Generation. In Noble Metal-Free Electrocatalysts: New Trends in Electrocatalysts for Energy Applications; American Chemical Society: Washington, DC, USA, 2022; pp. 191–235. [Google Scholar]
  40. Meng, S.; An, P.; Chen, L.; Sun, S.; Xie, Z.; Chen, M.; Jiang, D. Integrating Ru-modulated CoP nanosheets binary co-catalyst with 2D g-C3N4 nanosheets for enhanced photocatalytic hydrogen evolution activity. J. Colloid Interface Sci. 2021, 585, 108–117. [Google Scholar] [CrossRef] [PubMed]
  41. Cao, S.; Wang, C.J.; Fu, W.F.; Chen, Y. Metal phosphides as co-catalysts for photocatalytic and photoelectrocatalytic water splitting. ChemSusChem 2017, 10, 4306–4323. [Google Scholar] [CrossRef] [Green Version]
  42. Theerthagiri, J.; Murthy, A.P.; Lee, S.J.; Karuppasamy, K.; Arumugam, S.R.; Yu, Y.; Hanafiah, M.M.; Kim, H.S.; Mittal, V.; Choi, M.Y. Recent progress on synthetic strategies and applications of transition metal phosphides in energy storage and conversion. Ceram. Int. 2021, 47, 4404–4425. [Google Scholar] [CrossRef]
  43. Fang, Y.; Flake, J.C. Electrochemical reduction of CO2 at functionalized Au electrodes. J. Am. Chem. Soc. 2017, 139, 3399–3405. [Google Scholar] [CrossRef]
  44. Antil, B.; Kumar, L.; Das, M.R.; Deka, S. Incorporating NiCoP Cocatalyst into Hollow Rings of ZnCo-Metal–Organic Frameworks to Deliver Pt Cocatalyst like Visible Light Driven Hydrogen Evolution Activity. ACS Appl. Energy Mater. 2022, 5, 11113–11121. [Google Scholar] [CrossRef]
  45. Koshy, D.M.; Akhade, S.A.; Shugar, A.; Abiose, K.; Shi, J.; Liang, S.; Oakdale, J.S.; Weitzner, S.E.; Varley, J.B.; Duoss, E.B.; et al. Chemical modifications of Ag catalyst surfaces with imidazolium ionomers modulate H2 evolution rates during electrochemical CO2 reduction. J. Am. Chem. Soc. 2021, 143, 14712–14725. [Google Scholar] [CrossRef]
  46. Liu, Y.; Li, X.; He, H.; Yang, S.; Jia, G.; Liu, S. CoP imbedded g-C3N4 heterojunctions for highly efficient photo, electro and photoelectrochemical water splitting. J. Colloid Interface Sci. 2021, 599, 23–33. [Google Scholar] [CrossRef] [PubMed]
  47. Guo, F.; Huang, X.; Chen, Z.; Sun, H.; Chen, L. Prominent co-catalytic effect of CoP nanoparticles anchored on high-crystalline g-C3N4 nanosheets for enhanced visible-light photocatalytic degradation of tetracycline in wastewater. J. Chem. Eng. 2020, 395, 125118. [Google Scholar] [CrossRef]
  48. Sun, Z.; Zhu, M.; Fujitsuka, M.; Wang, A.; Shi, C.; Majima, T. Phase Effect of NixPy Hybridized with g-C3N4 for Photocatalytic Hydrogen Generation. ACS Appl. Mater. Interfaces 2017, 9, 30583–30590. [Google Scholar] [CrossRef] [PubMed]
  49. Sun, Y.; Wang, S.; Jiao, D.; Li, F.; Qiu, S.; Wang, Z.; Cai, Q.; Zhao, J.; Sun, C. Two-dimensional Pt2P3 monolayer: A promising bifunctional electrocatalyst with different active sites for hydrogen evolution and CO2 reduction. Chin. Chem. Lett. 2022, 33, 3987–3992. [Google Scholar] [CrossRef]
  50. Sun, M.; Liu, H.; Qu, J.; Li, J. Earth-rich transition metal phosphide for energy conversion and storage. Adv. Energy Mater. 2016, 6, 1600087. [Google Scholar] [CrossRef]
  51. Zhao, H.; Yuan, Z.Y. Insights into transition metal phosphate materials for efficient electrocatalysis. ChemCatChem 2020, 12, 3797–3810. [Google Scholar] [CrossRef]
  52. Jiao, Y.Q.; Yan, H.J.; Tian, C.G.; Fu, H.G. Structure Engineering and Electronic Modulation of Transition Metal Interstitial Compounds for Electrocatalytic Water Splitting. Acc. Chem. Res. 2022, 4, 42–56. [Google Scholar] [CrossRef]
  53. Huang, C.J.; Xu, H.M.; Shuai, T.Y.; Zhan, Q.N.; Zhang, Z.J.; Li, G.R. A review of modulation strategies for improving catalytic performance of transition metal phosphides for oxygen evolution reaction. Appl. Catal. B 2022, 325, 122313. [Google Scholar] [CrossRef]
  54. He, R.; Huang, X.; Feng, L. Recent progress in transition-metal sulfide catalyst regulation for improved oxygen evolution reaction. Energy Fuels 2022, 36, 6675–6694. [Google Scholar] [CrossRef]
  55. Oyama, S.T.; Gott, T.; Zhao, H.; Lee, Y.K. Transition metal phosphide hydro processing catalysts: A review. J. Catal. Today 2009, 143, 94–107. [Google Scholar] [CrossRef]
  56. Gong, N.; Deng, C.; Wu, L.; Wan, B.; Wang, Z.; Li, Z.; Gou, H.; Gao, F. Structural diversity and electronic properties of 3d transition metal tetraphosphides, TMP4 (TM= V, Cr, Mn, and Fe). Inorg. Chem. 2018, 57, 9385–9392. [Google Scholar] [CrossRef] [PubMed]
  57. Lakshmy, S.; Santhosh, S.; Kalarikkal, N.; Rout, C.S.; Chakraborthy, B. A review of electrochemical glucose sensing based on transition metal phosphides. J. Appl. Phys. 2023, 133, 070702. [Google Scholar] [CrossRef]
  58. Kuo, D.Y.; Nishiwaki, E.; Rivera-Maldonado, R.A.; Cossairt, B.M. The Role of Hydrogen Adsorption Site Diversity in Catalysis on Transition-Metal Phosphide Surfaces. ACS Catal. 2022, 13, 287–295. [Google Scholar] [CrossRef]
  59. Shi, Y.; Li, M.; Yu, Y.; Zhang, B. Recent advances in nanostructured transition metal phosphides: Synthesis and energy-related applications. Energy Environ. Sci. 2020, 13, 4564–4582. [Google Scholar] [CrossRef]
  60. Weng, B.; Wei, W.; Wu, H.; Alenizi, A.M.; Zheng, G. Bifunctional CoP and CoN porous nanocatalysts derived from ZIF-67 in situ grown on nanowire photoelectrodes for efficient photoelectrochemical water splitting and CO2 reduction. J. Mater. Chem. A 2016, 4, 15353–15360. [Google Scholar] [CrossRef]
  61. Lin, L.; Piao, S.; Choi, Y.; Lyu, L.; Hong, H.; Kim, D.; Lee, J.; Zhang, W.; Piao, Y. Nanostructured transition metal nitrides as emerging electrocatalysts for water electrolysis: Status and challenges. EnergyChem 2022, 4, 100072. [Google Scholar] [CrossRef]
  62. Zhong, Y.; Yin, L.; He, P.; Liu, W.; Wu, Z.; Wang, H. Surface chemistry in cobalt phosphide-stabilized lithium–sulfur batteries. J. Am. Chem. Soc. 2018, 140, 1455–1459. [Google Scholar] [CrossRef]
  63. Li, H.; Wen, P.; Itanze, D.S.; Hood, Z.D.; Ma, X.; Kim, M.; Adhikari, S.; Lu, C.; Dun, C.; Chi, M.; et al. Colloidal silver diphosphide (AgP2) nanocrystals as low overpotential catalysts for CO2 reduction to tunable syngas. Nat. Commun. 2019, 10, 5724. [Google Scholar] [CrossRef] [Green Version]
  64. Hu, E.; Feng, Y.; Nai, J.; Zhao, D.; Hu, Y.; Lou, X.W.D. Construction of hierarchical Ni–Co–P hollow nanobricks with oriented nanosheets for efficient overall water splitting. Energy Environ. Sci 2018, 11, 872–880. [Google Scholar] [CrossRef]
  65. Shi, Y.; Zhang, B. Engineering transition metal phosphide nanomaterials as highly active electrocatalysts for water splitting. J. Chem. Soc. Dalton Trans. 2017, 46, 16770–16773. [Google Scholar] [CrossRef] [PubMed]
  66. Xu, J.; Xiong, D.; Amorim, I.; Liu, L. Template-free synthesis of hollow iron phosphide–phosphate composite nanotubes for use as active and stable oxygen evolution electrocatalysts. ACS Appl. Nano Mater. 2018, 1, 617–624. [Google Scholar] [CrossRef]
  67. Yang, Y.; Luo, M.; Xing, Y.; Wang, S.; Zhang, W.; Lv, F.; Li, Y.; Zhang, Y.; Wang, W.; Guo, S. A Universal Strategy for Intimately Coupled Carbon Nanosheets/MoM Nanocrystals (M= P, S, C, and O) Hierarchical Hollow Nanospheres for Hydrogen Evolution Catalysis and Sodium-Ion Storage. J. Adv. Mater. 2018, 30, 1706085. [Google Scholar]
  68. Wu, T.; Pi, M.; Wang, X.; Guo, W.; Zhang, D.; Chen, S. Hierarchical cobalt poly-phosphide hollow spheres as highly active and stable electrocatalysts for hydrogen evolution over a wide pH range. Appl. Surf. Sci. 2018, 427, 800–806. [Google Scholar] [CrossRef]
  69. Yang, S.; Chen, G.; Ricciardulli, A.G.; Zhang, P.; Zhang, Z.; Shi, H.; Ma, J.; Zhang, J.; Blom, P.W.; Feng, X. Topochemical Synthesis of Two-Dimensional Transition-Metal Phosphides Using Phosphorene Templates. Angew. Chem. Int. Edi. 2020, 59, 465–470. [Google Scholar] [CrossRef]
  70. Zong, Q.; Liu, C.; Yang, H.; Zhang, Q.; Cao, G. Tailoring nanostructured transition metal phosphides for high-performance hybrid supercapacitors. Nano Today 2021, 38, 101201. [Google Scholar] [CrossRef]
  71. Su, J.; Zhou, J.; Wang, L.; Liu, C.; Chen, Y. Synthesis and application of transition metal phosphides as electrocatalyst for water splitting. Sci. Bull. 2017, 62, 633–644. [Google Scholar] [CrossRef] [Green Version]
  72. Li, Q.; Feng, W.; Liu, Y.; Chen, D.; Wu, Z.; Wang, H. Synergistic effect of spatially isolated Ni2P and NiO redox cocatalysts on g-C3N4 for sustainably boosted CO2 photocatalytic reduction. J. Mater. Chem. A 2022, 10, 15752–15765. [Google Scholar] [CrossRef]
  73. Liu, Z.; Yang, S.; Sun, B.; Chang, X.; Zheng, J.; Li, X. A peapod-like CoP@ C nanostructure from phosphorization in a low-temperature molten salt for high-performance lithium-ion batteries. Angew. Chem. Int. Ed. Engl. 2018, 57, 10187–10191. [Google Scholar] [CrossRef] [PubMed]
  74. Li, D.Y.; Liao, L.L.; Zhou, H.Q.; Zhao, Y.; Cai, F.M.; Zeng, J.S.; Liu, F.; Wu, H.; Tang, D.S.; Yu, F. Highly active non-noble electrocatalyst from Co2P/Ni2P nanohybrids for pH-universal hydrogen evolution reaction. Mater. Today Phys. 2021, 16, 100314. [Google Scholar] [CrossRef]
  75. Cho, G.; Park, Y.; Kang, H.; Hong, Y.K.; Lee, T.; Ha, D.H. Transition metal-doped FeP nanoparticles for hydrogen evolution reaction catalysis. Appl. Surf. Sci. 2020, 510, 145427. [Google Scholar] [CrossRef]
  76. Xiong, L.; Wang, B.; Cai, H.; Hao, H.; Li, J.; Yang, T.; Yang, S. Understanding the doping effect on hydrogen evolution activity of transition-metal phosphides: Modeled with Ni2P. Appl. Catal. B 2021, 295, 120283. [Google Scholar] [CrossRef]
  77. Ye, C.; Wang, M.Q.; Chen, G.; Deng, Y.H.; Li, L.J.; Luo, H.Q.; Li, N.B. One-step CVD synthesis of carbon framework wrapped Co2P as a flexible electrocatalyst for efficient hydrogen evolution. J. Mater. Chem. A 2017, 5, 7791–7795. [Google Scholar] [CrossRef]
  78. Li, X.L.; Wang, X.J.; Zhu, J.Y.; Li, Y.P.; Zhao, J.; Li, F.T. Fabrication of two-dimensional Ni2P/ZnIn2S4 heterostructures for enhanced photocatalytic hydrogen evolution. J. Chem. Eng. 2018, 353, 15–24. [Google Scholar] [CrossRef]
  79. Sun, W.; Fu, Z.; Shi, H.; Jin, C.; Liu, E.; Zhang, X.; Fan, J. Cu3P and Ni2P co-modified g-C3N4 nanosheet with excellent photocatalytic H2 evolution activities. J. Chem. Technol. Biotechnol. 2020, 95, 3117–3125. [Google Scholar] [CrossRef]
  80. Su, L.; Chen, M.; Zhang, H.; Tu, W. Construction of porous Ni2P cocatalyst and its promotion effect on photocatalytic H2 production reaction and CO2 reduction. Int. J. Hydrogen Energy 2023, 48, 15105–15116. [Google Scholar] [CrossRef]
  81. Fu, Z.C.; Xu, R.C.; Moore, J.T.; Liang, F.; Nie, X.C.; Mi, C.; Mo, J.; Xu, Y.; Xu, Q.Q.; Yang, Z.; et al. Highly Efficient Photocatalytic System Constructed from CoP/Carbon Nanotubes or Graphene for Visible-Light-Driven CO2 Reduction. Chem. Eur. J. 2018, 24, 4273–4278. [Google Scholar] [CrossRef]
  82. Ji, L.; Li, L.; Ji, X.; Zhang, Y.; Mou, S.; Wu, T.; Liu, Q.; Li, B.; Zhu, X.; Luo, Y.; et al. Highly selective electrochemical reduction of CO2 to alcohols on an FeP nanoarray. Angew. Chem. 2020, 132, 768–772. [Google Scholar] [CrossRef]
  83. Guo, Y.; Wang, Q.; Wang, M.; Shen, M.; Zhang, L.; Shi, J. FeP modified polymeric carbon nitride as a noble-metal-free photocatalyst for efficient CO2 reduction. Catal. Commun. 2021, 156, 106326. [Google Scholar] [CrossRef]
  84. Ren, Y.; Li, Z.; Deng, B.; Ye, C.; Zhang, L.; Wang, Y.; Li, T.; Liu, Q.; Cui, G.; Asiri, A.M.; et al. Superior hydrogen evolution electrocatalysis enabled by CoP nanowire array on graphite felt. Int. J. Hydrogen Energy 2022, 47, 3580–3586. [Google Scholar] [CrossRef]
  85. Jin, C.; Xu, C.; Chang, W.; Ma, X.; Hu, X.; Liu, E.; Fan, J. Bimetallic phosphide NiCoP anchored g-C3N4 nanosheets for efficient photocatalytic H2 evolution. J. Alloys Compd. 2019, 803, 205–215. [Google Scholar] [CrossRef]
  86. Gong, S.; Hou, M.; Niu, Y.; Teng, X.; Liu, X.; Xu, M.; Xu, C.; Au, V.K.M.; Chen, Z. Molybdenum phosphide coupled with highly dispersed nickel confined in porous carbon nanofibers for enhanced photocatalytic CO2 reduction. J. Chem. Eng. 2022, 427, 131717. [Google Scholar] [CrossRef]
  87. Zhu, J.; He, Q.; Liu, Y.; Key, J.; Nie, S.; Wu, M.; Shen, P.K. Three-dimensional, hetero-structured, Cu3P@C nanosheets with excellent cycling stability as Na-ion battery anode material. J. Mater. Chem. 2019, 7, 16999–17007. [Google Scholar] [CrossRef]
  88. Hong, L.F.; Guo, R.T.; Yuan, Y.; Ji, X.Y.; Lin, Z.D.; Li, Z.S.; Pan, W.G. Recent progress of transition metal phosphides for photocatalytic hydrogen evolution. ChemSusChem 2021, 14, 539–557. [Google Scholar] [CrossRef] [PubMed]
  89. Li, Y.H.; Qi, M.Y.; Li, J.Y.; Tang, Z.R.; Xu, Y.J. Noble metal free CdS@ CuS-NixP hybrid with modulated charge transfer for enhanced photocatalytic performance. Appl. Catal. B 2019, 257, 117934. [Google Scholar] [CrossRef]
  90. Zou, Y.; Guo, C.; Cao, X.; Zhang, L.; Chen, T.; Guo, C.; Wang, J. Synthesis of CdS/CoP hollow nanocages with improved photocatalytic water splitting performance for hydrogen evolution. J. Environ. Chem. Eng. 2021, 9, 106270. [Google Scholar] [CrossRef]
  91. Ray, A.; Sultana, S.; Paramanik, L.; Parida, K.M. Recent advances in phase, size, and morphology-oriented nanostructured nickel phosphide for overall water splitting. J. Mater. Chem. A 2020, 8, 19196–19245. [Google Scholar] [CrossRef]
  92. Ali, S.A.; Ahmad, T. Enhanced hydrogen generation via overall water splitting using novel MoS2-BN nanoflowers assembled TiO2 ternary heterostructures. Int. J. Hydrog. Energy 2023, 48, 22044–22059. [Google Scholar] [CrossRef]
  93. Zhang, X.; Yan, J.; Zheng, F.; Zhao, J.; Lee, L.Y.S. Designing charge transfer route at the interface between WP nanoparticle and g-C3N4 for highly enhanced photocatalytic CO2 reduction reaction. Appl. Catal. B 2021, 286, 119879. [Google Scholar] [CrossRef]
  94. Mehtab, A.; Banerjee, S.; Mao, Y.; Ahmad, T. Type-II CuFe2O4/graphitic carbon nitride heterojunctions for high-efficiency photocatalytic and electrocatalytic hydrogen generation. ACS Appl. Mater. Interfaces 2022, 14, 44317–44329. [Google Scholar] [CrossRef] [PubMed]
  95. Shen, R.; Xie, J.; Ding, Y.; Liu, S.Y.; Adamski, A.; Chen, X.; Li, X. Carbon nanotube-supported Cu3P as high-efficiency and low-cost cocatalysts for exceptional semiconductor-free photocatalytic H2 evolution. ACS Sustain. Chem. Eng. 2018, 7, 3243–3250. [Google Scholar] [CrossRef]
  96. Song, T.; Zhang, X.; Matras-Postolek, K.; Yang, P. N-doped carbon layer promoted charge separation/transfer in WP/g-C3N4 heterostructures for efficient H2 evolution and 4-nitrophenol removal. Carbon 2023, 202, 378–388. [Google Scholar] [CrossRef]
  97. Wang, F.; Qian, G.; Kong, X.P.; Zhao, X.; Hou, T.; Chen, L.; Fang, R.; Li, Y. Hierarchical double-shelled CoP nanocages for efficient visible-light-driven CO2 reduction. ACS Appl. Mater. Interfaces 2021, 13, 45609–45618. [Google Scholar] [CrossRef] [PubMed]
  98. Lv, X.; Li, X.; Yang, C.; Ding, X.; Zhang, Y.; Zheng, Y.Z.; Li, S.; Sun, X.; Tao, X. Large-size, porous, ultrathin NiCoP nanosheets for efficient electro/photocatalytic water splitting. Adv. Funct. Mater. 2020, 30, 1910830. [Google Scholar] [CrossRef]
  99. Dou, M.Y.; Han, S.R.; Du, X.X.; Pang, D.H.; Li, L.L. Well-defined FeP/CdS heterostructure construction with the assistance of amine for the efficient H2 evolution under visible light irradiation. Int. J. Hydrogen Energy 2020, 45, 32039–32049. [Google Scholar] [CrossRef]
  100. Mishra, I.K.; Zhou, H.Q.; Sun, J.Y.; Qin, F.; Dahal, K.; Bao, J.M.; Chen, S.; Ren, Z.F. Hierarchical CoP/Ni5P4/CoP microsheet arrays as a robust pH-universal electrocatalyst for efficient hydrogen generation. Energy Environ. Sci. 2018, 11, 2246–2252. [Google Scholar] [CrossRef]
  101. Tian, J.Q.; Liu, Q.; Asiri, A.M.; Sun, X.P. Self-supported nanoporous cobalt phosphide nanowire arrays: An efficient 3D hydrogen-evolving cathode over the wide range of pH 0-14. J. Am. Chem. Soc. 2014, 136, 7587–7590. [Google Scholar] [CrossRef]
  102. Han, A.L.; Chen, H.L.; Zhang, H.Y.; Sun, Z.J.; Du, P.W. Ternary metal phosphide nanosheets as a highly efficient electrocatalyst for water reduction to hydrogen over a wide pH range from 0 to 14. J. Mater. Chem. A 2016, 4, 10195–10202. [Google Scholar] [CrossRef]
  103. Li, J.; Zheng, H.; Xu, C.; Su, Z.; Li, X.; Sun, J. Bimetallic phosphides as high-efficient electrocatalysts for hydrogen generation. Inorg. Chem. 2021, 60, 1624–1630. [Google Scholar] [CrossRef]
  104. Yang, M.; Jiang, Y.; Qu, M.; Qin, Y.; Wang, Y.; Shen, W.; He, R.; Su, W.; Li, M. Strong electronic couple engineering of transition metal phosphides-oxides heterostructures as multifunctional electrocatalyst for hydrogen production. Appl. Catal. B Environ. 2020, 269, 118803. [Google Scholar] [CrossRef]
  105. Kang, Q.; Li, M.; Shi, J.; Lu, Q.; Gao, F. A universal strategy for carbon-supported transition metal phosphides as high-performance bifunctional electrocatalysts towards efficient overall water splitting. ACS Appl. Mater. Interfaces 2020, 12, 19447–19456. [Google Scholar] [CrossRef]
  106. Chen, D.; Pu, Z.; Lu, R.; Ji, P.; Wang, P.; Zhu, J.; Lin, C.; Li, H.W.; Zhou, X.; Hu, Z.; et al. Ultralow Ru loading transition metal phosphides as high-efficient bifunctional electrocatalyst for a solar-to-hydrogen generation system. Adv. Energy Mater. 2020, 10, 2000814. [Google Scholar] [CrossRef]
  107. Liu, Y.; Zhang, Z.; Zhang, L.; Xia, Y.; Wang, H.; Liu, H.; Ge, S.; Yu, J. Manipulating the d-band centers of transition metal phosphides through dual metal doping towards robust overall water splitting. J. Mater. Chem. A 2022, 10, 22125–22134. [Google Scholar] [CrossRef]
  108. Zhang, H.; Aierke, A.; Zhou, Y.; Ni, Z.; Feng, L.; Chen, A.; Wågberg, T.; Hu, G. A High-performance transition-metal phosphide electrocatalyst for converting solar energy into hydrogen at 19.6% STH efficiency. Carbon Energy 2023, 5, 217. [Google Scholar] [CrossRef]
  109. Pei, Y.; Cheng, Y.; Chen, J.; Smith, W.; Dong, P.; Ajayan, P.M.; Ye, M.; Shen, J. Recent developments of transition metal phosphides as catalysts in the energy conversion field. J. Mater. Chem. A 2018, 6, 23220–23243. [Google Scholar] [CrossRef]
  110. Li, Y.; Dong, Z.; Jiao, L. Multifunctional Transition Metal-Based Phosphides in Energy-Related Electrocatalysis. Adv. Energy Mater. 2020, 10, 1902104. [Google Scholar] [CrossRef]
  111. Downes, C.A.; Libretto, N.J.; Harman-Ware, A.E.; Happs, R.M.; Ruddy, D.A.; Baddour, F.G.; Ferrell III, J.R.; Habas, S.E.; Schaidle, J.A. Electrocatalytic CO2 reduction over Cu3P nanoparticles generated via a molecular precursor route. ACS Appl. Energy Mater. 2020, 3, 10435–10446. [Google Scholar] [CrossRef]
  112. Sun, X.; Lu, L.; Zhu, Q.; Wu, C.; Yang, D.; Chen, C.; Han, B. MoP nanoparticles supported on indium-doped porous carbon: Outstanding catalysts for highly efficient CO2 electroreduction. Angew. Chem. Int. Ed. Eng. 2018, 57, 2427–2431. [Google Scholar] [CrossRef]
  113. Kim, M.G.; Choi, Y.; Park, E.; Cheon, C.H.; Kim, N.K.; Min, B.K.; Kim, W. Crystalline/Amorphous Ni2P/Ho2O3 Core/Shell Nanoparticles for Electrochemical Reduction of CO2 to Acetone. ACS Appl. Energy Mater. 2020, 3, 11516–11522. [Google Scholar] [CrossRef]
  114. Laursen, A.B.; Calvinho, K.U.; Goetjen, T.A.; Yap, K.M.; Hwang, S.; Yang, H.; Garfunkel, E.; Dismukes, G.C. CO2 electro-reduction on Cu3P: Role of Cu (I) oxidation state and surface facet structure in C1-formate production and H2 selectivity. Electrochim. Acta 2021, 391, 138889. [Google Scholar] [CrossRef]
  115. Calvinho, K.U.; Alherz, A.W.; Yap, K.M.; Laursen, A.B.; Hwang, S.; Bare, Z.J.; Clifford, Z.; Musgrave, C.B.; Dismukes, G.C. Surface Hydrides on Fe2P Electrocatalyst Reduce CO2 at Low Overpotential: Steering Selectivity to Ethylene Glycol. J. Am. Chem. Soc. 2021, 143, 21275–21285. [Google Scholar] [CrossRef] [PubMed]
  116. Banerjee, S.; Kakekhani, A.; Wexler, R.B.; Rappe, A.M. Mechanistic Insights into CO2 Electroreduction on Ni2P: Understanding Its Selectivity toward Multicarbon Products. ACS Catal. 2021, 11, 11706–11715. [Google Scholar] [CrossRef]
  117. Rosser, T.E.; Windle, C.D.; Reisner, E. Electrocatalytic and Solar-Driven CO2 Reduction to CO with a Molecular Manganese Catalyst Immobilized on Mesoporous TiO2. Angew. Chem. Int. Ed. Eng 2016, 55, 7388–7392. [Google Scholar]
  118. Liu, T.; Li, P.; Yao, N.; Cheng, G.Z.; Chen, S.L.; Luo, W.; Yin, Y.D. CoP-doped MOF-based electrocatalyst for pH-universal hydrogen evolution reaction. Angew. Chem. Int. Ed. 2019, 58, 4679–4684. [Google Scholar] [CrossRef]
  119. Pan, Y.; Liu, Y.; Liu, C. Nanostructured nickel phosphide supported on carbon nanospheres: Synthesis and application as an efficient electrocatalyst for hydrogen evolution. J. Power Sources 2015, 285, 169–177. [Google Scholar] [CrossRef]
  120. Niu, P.; Pan, Z.; Wang, S.; Wang, X. Cobalt Phosphide Cocatalysts Coated with Porous N-doped Carbon Layers for Photocatalytic CO2 Reduction. ChemCatChem 2021, 13, 3581–3587. [Google Scholar] [CrossRef]
  121. Zhang, Z.; Song, N.; Wang, J.; Liu, Y.; Dai, Z.; Nie, G. Polydopamine-derived carbon layer anchoring NiCo-P nanowire arrays for high-performance binder-free supercapacitor and electrocatalytic hydrogen evolution. SusMat 2022, 2, 646–657. [Google Scholar] [CrossRef]
  122. Wang, J.; Liu, Z.; Zheng, Y.; Cui, L.; Yang, W.; Liu, J. Recent advances in cobalt phosphide-based materials for energy-related applications. J. Mater. Chem. A 2017, 5, 22913–22932. [Google Scholar]
  123. Shah, S.S.A.; Khan, N.A.; Imran, M.; Rashid, M.; Tufail, M.K.; Rehman, A.U.; Balkourani, G.; Sohail, M.; Najam, T.; Tsiakaras, P. Recent Advances in Transition Metal Tellurides (TMTs) and Phosphides (TMPs) for Hydrogen Evolution Electrocatalysis. Membranes 2023, 13, 113. [Google Scholar] [CrossRef] [PubMed]
  124. Zhao, H.; Yuan, Z.Y. Transition metal–phosphorus-based materials for electrocatalytic energy conversion reactions. Catal. Sci. Technol. 2017, 7, 330–347. [Google Scholar] [CrossRef]
  125. Chu, Y.; Wang, D.; Shan, X.; Liu, C.; Wang, W.; Mitsuzaki, N.; Chen, Z. Activity engineering to transition metal phosphides as bifunctional electrocatalysts for efficient water-splitting. Int. J. Hydrogen Energy 2022, 47, 38983–39000. [Google Scholar] [CrossRef]
  126. Du, H.; Kong, R.M.; Guo, X.; Qu, F.; Li, J. Recent progress in transition metal phosphides with enhanced electrocatalysis for hydrogen evolution. Nanoscale 2018, 10, 21617–21624. [Google Scholar] [CrossRef] [PubMed]
  127. Zeng, D.; Xu, W.; Ong, W.J.; Xu, J.; Ren, H.; Chen, Y.; Zheng, H.; Peng, D.L. Toward noble-metal-free visible-light-driven photocatalytic hydrogen evolution: Monodisperse sub–15 nm Ni2P nanoparticles anchored on porous g-C3N4 nanosheets to engineer 0D-2D heterojunction interfaces. Appl. Catal. B 2018, 221, 47–55. [Google Scholar] [CrossRef]
  128. Zhang, R.; Wang, X.; Yu, S.; Wen, T.; Zhu, X.; Yang, F.; Sun, X.; Wang, X.; Hu, W. Ternary NiCo2Px nanowires as pH-universal electrocatalysts for highly efficient hydrogen evolution reaction. Adv. Mater. 2017, 29, 1605502. [Google Scholar] [CrossRef] [PubMed]
  129. Downes, C.A.; Van Allsburg, K.M.; Tacey, S.A.; Unocic, K.A.; Baddour, F.G.; Ruddy, D.A.; LiBretto, N.J.; O’Connor, M.M.; Farberow, C.A.; Schaidle, J.A.; et al. Controlled Synthesis of Transition Metal Phosphide Nanoparticles to Establish Composition-Dependent Trends in Electrocatalytic Activity. Chem. Mater. 2022, 34, 6255–6267. [Google Scholar] [CrossRef]
  130. Irshad, S.; Ullah, F.; Khan, S.; Ludwig, R.; Mahmood, T.; Ayub, K. First row transition metals decorated boron phosphide nanoclusters as nonlinear optical materials with high thermodynamic stability and enhanced electronic properties; A detailed quantum chemical study. Opt. Laser Technol. 2021, 134, 106570. [Google Scholar] [CrossRef]
Figure 1. Schematic demonstration of photocatalytic reduction/conversion of CO2 into solar fuel. [Reprinted with permission from Ref. [1]. Copyright 2023, John Wiley and Sons].
Figure 1. Schematic demonstration of photocatalytic reduction/conversion of CO2 into solar fuel. [Reprinted with permission from Ref. [1]. Copyright 2023, John Wiley and Sons].
Catalysts 13 01046 g001
Figure 2. Diagrammatic illustration of CO2RR and HER with respect to noble metal and transition metal phosphide.
Figure 2. Diagrammatic illustration of CO2RR and HER with respect to noble metal and transition metal phosphide.
Catalysts 13 01046 g002
Figure 3. Schematic landscape of TMPs synthesis, properties, and energy driven applications.
Figure 3. Schematic landscape of TMPs synthesis, properties, and energy driven applications.
Catalysts 13 01046 g003
Figure 4. (a) Triangular prism and tetrakaidecahedral structures in phosphides and (b) crystal structures of metal-rich phosphides. [Reprinted with permission from Ref. [55]. Copyright 2009, Catalysis].
Figure 4. (a) Triangular prism and tetrakaidecahedral structures in phosphides and (b) crystal structures of metal-rich phosphides. [Reprinted with permission from Ref. [55]. Copyright 2009, Catalysis].
Catalysts 13 01046 g004
Figure 5. Crystal structures of 3D-TMP4 (TM = V, Cr, Mn, and Fe) polyhedron view. The TM and P atoms are represented as big yellow spheres and small green spheres, respectively. [Reprinted with permission from Ref. [56]. Copyright 2018, American Chemical Society].
Figure 5. Crystal structures of 3D-TMP4 (TM = V, Cr, Mn, and Fe) polyhedron view. The TM and P atoms are represented as big yellow spheres and small green spheres, respectively. [Reprinted with permission from Ref. [56]. Copyright 2018, American Chemical Society].
Catalysts 13 01046 g005
Figure 6. (a) The mechanism of Ni-MoP@NCPF as a photocatalyst for CO2 reduction, (b) CO generation on various photocatalysts under visible light irradiation (with a UVCUT-420 nm filter) for 3 h, and (c) gas products of CO2 reduction on Ni-MoP@NCPF under visible light irradiation for 3 h. [Reprinted with permission from Ref. [86]. Copyright 2022, Elsevier].
Figure 6. (a) The mechanism of Ni-MoP@NCPF as a photocatalyst for CO2 reduction, (b) CO generation on various photocatalysts under visible light irradiation (with a UVCUT-420 nm filter) for 3 h, and (c) gas products of CO2 reduction on Ni-MoP@NCPF under visible light irradiation for 3 h. [Reprinted with permission from Ref. [86]. Copyright 2022, Elsevier].
Catalysts 13 01046 g006
Figure 7. Mechanism of photocatalytic H2 production for Cu3P-Ni2P/g-C3N4. [Reprinted with permission from Ref. [79]. Copyright 2020, John Wiley and Sons].
Figure 7. Mechanism of photocatalytic H2 production for Cu3P-Ni2P/g-C3N4. [Reprinted with permission from Ref. [79]. Copyright 2020, John Wiley and Sons].
Catalysts 13 01046 g007
Figure 8. Photocatalytic rates of CO2 reduction over porous and non-porous Ni2P/CdS composites: (a) CO, (b) CH4, (c) H2, (d) survey of various gases. [Reprinted with permission from Ref. [80]. Copyright 2023, Elsevier].
Figure 8. Photocatalytic rates of CO2 reduction over porous and non-porous Ni2P/CdS composites: (a) CO, (b) CH4, (c) H2, (d) survey of various gases. [Reprinted with permission from Ref. [80]. Copyright 2023, Elsevier].
Catalysts 13 01046 g008
Figure 9. (a) LSVs of PC, NiP@C, MoP@C, MoNiP@C, MoP/MoNiP, MoP/MoNiP@C, and Pt/C in 0.5 M H2SO4 solution and (b) Tafel plots of PC, NiP@C, MoP@C, MoNiP@C, MoP/MoNiP, MoP/MoNiP@C, and Pt/C. (c) CV plots of MoP/MoNiP@C at a scan rates ranging from 25 to 250 mV s−1 and (d) linear fitting of Δj vs. scan rates of MoP/MoNiP@C. (e) EIS spectra of PC, NiP@C, MoP@CMoNiP@C, MoP/MoNiP, and MoP/MoNiP@C. (f) LSVs of MoP/MoNiP@C in 0.5 M H2SO4 before and after 1000 cycles, and the inset indicates the time-dependent current density curve at an overpotential of 140 mV for 24 h. [Reprinted with permission from Ref. [103]. Copyright 2021, American Chemical Society].
Figure 9. (a) LSVs of PC, NiP@C, MoP@C, MoNiP@C, MoP/MoNiP, MoP/MoNiP@C, and Pt/C in 0.5 M H2SO4 solution and (b) Tafel plots of PC, NiP@C, MoP@C, MoNiP@C, MoP/MoNiP, MoP/MoNiP@C, and Pt/C. (c) CV plots of MoP/MoNiP@C at a scan rates ranging from 25 to 250 mV s−1 and (d) linear fitting of Δj vs. scan rates of MoP/MoNiP@C. (e) EIS spectra of PC, NiP@C, MoP@CMoNiP@C, MoP/MoNiP, and MoP/MoNiP@C. (f) LSVs of MoP/MoNiP@C in 0.5 M H2SO4 before and after 1000 cycles, and the inset indicates the time-dependent current density curve at an overpotential of 140 mV for 24 h. [Reprinted with permission from Ref. [103]. Copyright 2021, American Chemical Society].
Catalysts 13 01046 g009
Figure 10. Synthesis Process for MoP/MoNiP@C. [Reprinted with permission from Ref. [103]. Copyright 2021, American Chemical Society].
Figure 10. Synthesis Process for MoP/MoNiP@C. [Reprinted with permission from Ref. [103]. Copyright 2021, American Chemical Society].
Catalysts 13 01046 g010
Figure 11. Schematic illustration of the experimental procedure for synthesis of C/A Ni2P/Ho2O3 CSNPs (NPs, NDs, and CSNPs stand for nanoparticles, nanodisks, and core–shell nanoparticles, respectively). [Reprinted with permission from Ref. [113]. Copyright 2020, American Chemical Society].
Figure 11. Schematic illustration of the experimental procedure for synthesis of C/A Ni2P/Ho2O3 CSNPs (NPs, NDs, and CSNPs stand for nanoparticles, nanodisks, and core–shell nanoparticles, respectively). [Reprinted with permission from Ref. [113]. Copyright 2020, American Chemical Society].
Catalysts 13 01046 g011
Figure 12. Potential-dependent FEs of CO2RR products and total current densities obtained using (a) Ni2P/Ho2O3 CSNPs, (b) Ho2O3 NDs, and (c) Ni2P NPs as electrocatalysts. (d) Suggested CO2RR pathways toward (CH3)2CO. [Reprinted with permission from Ref. [113]. Copyright 2020, American Chemical Society].
Figure 12. Potential-dependent FEs of CO2RR products and total current densities obtained using (a) Ni2P/Ho2O3 CSNPs, (b) Ho2O3 NDs, and (c) Ni2P NPs as electrocatalysts. (d) Suggested CO2RR pathways toward (CH3)2CO. [Reprinted with permission from Ref. [113]. Copyright 2020, American Chemical Society].
Catalysts 13 01046 g012
Table 1. Transition metal-phosphide-based photocatalysts for HER and CO2RR applications.
Table 1. Transition metal-phosphide-based photocatalysts for HER and CO2RR applications.
CatalystsSynthesis MethodApplicationBand Gap (eV)Rate of HER/CO2RRRef.
Ru-CoP-1:8/GCNChemical reductionHER2.251172.5 μmol g−1 h−1[40]
Ni2P/NiO/CNIn situ gas–solid reactionCO2RR2.6CO (1.506 μmolg−1 h−1)
CH4 (0.29 μmolg−1 h−1)
[72]
Ni2P/ZnIn2S4HydrothermalHER~22066 μmol g−1 h−1[78]
Cu3P-Ni2P/g-C3N4SolvothermalHER2.76529.8 μmol g−1 h−1[79]
Ni2P/CdSHydrothermalHER and CO2RR2.16H2 (111.3 mmol g−1 h−1)
CO (178.0 μmol g−1 h−1)
CH4 (61.2 μmol g−1 h−1)
[80]
CoP/rGOHydrothermalCO2RR-CO (47,330 μmol g−1 h−1)[81]
CoP/CNTHydrothermalCO2RR-CO (39,510 μmol g−1 h−1)[81]
FeP/CNThermal decompositionCO2RR2.40CO (5.19 μmol g−1 h−1)[83]
NiCoP/g-C3N4Thermal polymerizationHER2.695162 μmol g−1 h−1[85]
Ni-MoP@NPPFElectrospinningCO2RR1.42CO (953.53 μmol g−1 h−1)[86]
WP-NC/g-C3N4Facile sonicationCO2RR2.8CO (376 μmol g−1h−1)[93]
WP-NC/g-C3N4Thermal polymerizationHER~2.81217.6 μmol g−1 h−1[96]
Fe doped CoPSelf-assemblyCO2RR-CO (21.0 μmol h−1)[97]
NiCoP nanosheetsWet chemical and phosphorizationHER-238.2 mmol g−1 h−1[98]
FeP/CdSSolvothermalHER2.3237.92 mmol g−1 h−1[99]
Table 2. Transition-metal-phosphide-based electrocatalysts for HER application.
Table 2. Transition-metal-phosphide-based electrocatalysts for HER application.
ElectrocatalystSynthesis MethodApplicationOverpotential
(mV)
Tafel Slope
(mV dec−1)
Ref.
Co2P/Ni2PThermal phosphorizationHER51-[74]
Co-FePPhosphorizationHER12663.6[75]
Co-Ni2PSynthetic methodHER3147[76]
MoP/MoNiP@CCalcination and phosphorizationHER13466[103]
Ni2P/MoO2/NFPhosphorizationHER3445.8[104]
NiFeP@CCalcinationHER16075.8[105]
Ru-MnFeP/NFPhosphorizationHER3569[106]
Ni-Mn-FePPhosphorizationHER103-[107]
Er-NiCoP/NFPhosphorizationHER46-[108]
Table 3. Transition-metal-phosphide-based electrocatalysts for CO2RR application.
Table 3. Transition-metal-phosphide-based electrocatalysts for CO2RR application.
ElectrocatalystsSynthesis MethodPower Density/
Current Density
Faradaic EfficiencyRef.
Cu3P/CHydrothermal2.6 mW cm−247% (CO)[28]
AgP2Self-assembly−8.7 mA cm−282.0% (CO)[63]
FeP/TMHydrothermal-94.3%
(CH3OH + C2H5OH)
[82]
Cu3PThermal decomposition-8% (Formate)[111]
MoP@In-PCSolid state43.8 mA cm−296.5% (HCOOH)[112]
Ni2P/Ho2O3Phosphorization0.95 mA cm−225.4% (Acetone)[113]
Cu3P NS/CuSelf-assembly-1.1% ± 0.6%
(Formic acid)
[114]
TiO2/MnPAnnealing-67% ± 5% (CO)
12.4% ± 1.4% (H2)
[117]
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

Shaheen, S.; Ali, S.A.; Mir, U.F.; Sadiq, I.; Ahmad, T. Recent Advances in Transition Metal Phosphide Nanocatalysts for H2 Evolution and CO2 Reduction. Catalysts 2023, 13, 1046. https://doi.org/10.3390/catal13071046

AMA Style

Shaheen S, Ali SA, Mir UF, Sadiq I, Ahmad T. Recent Advances in Transition Metal Phosphide Nanocatalysts for H2 Evolution and CO2 Reduction. Catalysts. 2023; 13(7):1046. https://doi.org/10.3390/catal13071046

Chicago/Turabian Style

Shaheen, Saman, Syed Asim Ali, Umar Farooq Mir, Iqra Sadiq, and Tokeer Ahmad. 2023. "Recent Advances in Transition Metal Phosphide Nanocatalysts for H2 Evolution and CO2 Reduction" Catalysts 13, no. 7: 1046. https://doi.org/10.3390/catal13071046

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