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

Modulation of the photocatalytic activity of MAPbI3 crystals via Nb2CTx MXenes for high-efficiency photocatalytic hydrogen production

Ruiyan Suna, Lili Gao*a, Deng Lib, Hua Wangc, Fan Yanga, Jin Wangc, Ke Haoa, Haijiao Xied and Ping Hu*a
aSchool of Metallurgy Engineering, Xi'an International Science and Technology Cooperation Base for Manufacturing of Special Powder and Powder Metallurgy, Xi'an University of Architecture and Technology, Xi'an 710055, China. E-mail: gll0706@xauat.edu.cn; huping@xauat.edu.cn
bSchool of Materials Science and Engineering, Shaanxi Normal University, Xi'an 710119, China
cSchool of Environmental and Municipal Engineering, Xi'an University of Architecture and Technology, Xi'an 710055, China
dHangzhou Yanqu Information Technology Co., Ltd., Hangzhou 310003, China

Received 28th July 2025 , Accepted 2nd September 2025

First published on 2nd September 2025


Abstract

This study addresses rapid charge recombination and instability in MAPbI3 photocatalysts for hydrogen production by constructing a Nb2CTx MXenes-modulated composite. Leveraging Nb2CTx metallic conductivity for efficient electron extraction and proton reduction sites, the composite enables robust photocatalytic HI splitting in strong acid. In situ coupling achieved intimate heterointerface contact with Z-scheme characteristics, yielding a remarkable hydrogen evolution rate of 12[thin space (1/6-em)]046.77 μmol h−1 g−1—a 344-fold enhancement over pristine MAPbI3—while retaining ∼85% activity after 5 cycles. UPS and theoretical calculations confirm a strong built-in electric field at the heterointerface accelerates carrier separation. Critically, the Z-scheme simultaneously suppresses recombination and preserves strong redox capabilities. This interface engineering synergistically enhances efficiency and stability, resolving the charge separation-redox capability trade-off in conventional type II heterojunctions.


1. Introduction

Hydrogen energy has been widely recognized as the most promising next-generation clean energy carrier because its mass-specific energy density is approximately three times greater than that of gasoline and zero-pollution combustion byproducts.1 As a crucial pathway for solar-to-chemical energy conversion, photocatalytic water splitting technology for hydrogen production relies fundamentally on the development of highly efficient and stable photocatalytic systems.2–8 Compared with conventional electrolytic water splitting, hydroiodic acid (HI) photolysis has significant advantages: an approximately 40% reduction in the reaction activation energy, a theoretical hydrogen production efficiency of 28%, and recyclable reactants under mild reaction conditions.9–14 High-performance photocatalysts must possess appropriate electronic structures capable of controlling the fate of photogenerated carriers to effectively suppress charge recombination.15–17 While traditional wide-bandgap semiconductors such as TiO2 (Eg = 3.2 eV) exhibit good chemical stability, their limited photoresponse range (<387 nm, UV region) and insufficient carrier diffusion length severely constrain solar utilization efficiency and charge separation18–20 capabilities. In recent years, organic–inorganic hybrid perovskite materials (ABX3-type) have attracted intensive research interest because of their unique photophysical properties. For example, methylammonium lead iodide (MAPbI3) possesses an optimal bandgap structure of 1.50 eV, enabling full-spectrum visible light absorption (absorption coefficient > 104 cm−1) and extended carrier lifetimes.21–26 Pioneering work by Park et al.13 first extended MAPbI3 applications to HI photolysis systems, establishing a new research direction in perovskite photocatalysis. However, MAPbI3 still faces two critical challenges in photocatalytic processes: rapid recombination of photogenerated carriers at the nanoscale, leading to reduced hydrogen evolution reaction (HER) rates, and structural instability-induced catalyst deactivation in HI solutions.11 These bottlenecks fundamentally originate from the intricate interplay between intrinsic electronic structure modulation and surface/interface stability optimization, urgently requiring breakthroughs through innovative material design and mechanistic understanding.

To address the aforementioned challenges, the construction of heterojunction interfaces has been recognized as a pivotal strategy for overcoming the performance bottleneck in perovskite photocatalysis by modulating carrier dynamics. By loading cocatalysts on MAPbI3 surfaces to form type II or Schottky heterojunctions,27 directional charge transport channels can be effectively established, thereby suppressing the recombination of photogenerated electron–hole pairs. Typical fabrication approaches involve depositing electron transport materials (e.g., reduced graphene oxide (rGO),28 few-layer black phosphorus (BP),14 Ni3C,10 MoS2,29,30 or transition metal compounds) as nanoparticles onto MAPbI3 perovskite surfaces. Noble metals such as Au, Pt, and Ag have also been widely employed as cocatalysts in heterojunction photocatalysts. For example, in the Pt–TiO2/MAPbI3 system, Pt nanoparticles function as electron traps that significantly enhance the carrier separation efficiency, achieving a hydrogen evolution rate of 1.99 mmol g−1 h−1 with a corresponding solar-to-hydrogen (STH) conversion efficiency of 0.86%.11 This performance enhancement originates from heterojunction-induced band bending, which drives the migration of photogenerated electrons from MAPbI3 to cocatalysts, thereby optimizing the kinetics of H+ reduction and substantially improving the photocatalytic hydrogen evolution reaction (HER) activity. Nevertheless, despite their effectiveness, noble metal cocatalysts (e.g., Pt and Au) face critical limitations in large-scale applications because of their low natural abundance and prohibitively high costs.31

In recent years, two-dimensional transition metal carbides/nitrides (MXenes) have emerged as novel candidates for developing non-noble metal cocatalysts because of their unique electronic configurations and surface characteristics.32–38 MXenes (with the general formula Mn+1XnTx) exhibit metal-like electrical conductivity (>104 S m−1), tunable surface functional groups (–O, –F, and –OH),39 and abundant exposed metallic sites. Compared with conventional carbon-based materials, their terminal metal sites demonstrate enhanced redox activity.40,41 Theoretical calculations revealed that heterointerfaces in MXenes-based composites can establish efficient charge transfer pathways, significantly improving the photogenerated charge carrier separation efficiency.42 The key mechanism lies in the Schottky barrier formed at the MXenes/semiconductor interface, which directionally drives photogenerated electron migration to the MXenes layer.43,44 Synchrotron radiation-based photoelectron spectroscopy studies confirm that charge accumulation induces a negative shift in the Fermi level of MXenes, with this band structure realignment substantially optimizing the thermodynamic driving force for photocatalytic hydrogen evolution reactions.45 Notably, the hydrogen adsorption Gibbs free energy (ΔGH*) of MXenes can be regulated through surface termination groups to approach the ideal value (ΔGH* ≈ 0) for optimal HER performance, as exemplified by hydrogenated W2C, oxidized Mo2C,46 O-terminated Ti3C2 (ref. 40) and other systems.47 Among various MXenes materials, Nb2C MXenes have distinctive advantages: their work function (∼4.3 eV) is highly compatible with that of common semiconductors (e.g., TiO2 and g-C3N4). The two-dimensional layered structure not only suppresses nanoparticle aggregation and enhances mechanical stability but also facilitates accelerated interfacial charge transfer through intrinsic carrier mobility at heterojunction interfaces, effectively reducing electron–hole recombination rates.48–51

Through interface engineering strategies, multiple efficient photocatalytic systems have been successfully constructed using Nb2CTx MXenes composites. Tayyab et al.52 developed an In2S3/Nb2O5/Nb2C ternary composite via in situ chemical anchoring, where Nb2C suppresses charge recombination while accelerating electron transfer from Nb2O5 and In2S3 photocatalysts, achieving a hydrogen evolution rate of 68.8 μmol g−1 h−1. In another study,53 Tayyab et al. established an electrostatic self-assembly system of CdS nanowires on Nb2C cocatalysts, wherein protons capture photoexcited electrons accumulated on Nb2C for efficient H2 generation. Su et al.54 reported a Nb2O5/C/Nb2C composite whose fourfold enhancement in photocatalytic water splitting over that of pure Nb2O5 originates from intimate Nb2O5–Nb2C interfacial contact and efficient carrier separation. Moreover, Xu's team55 employed a hydrothermal oxidation strategy to simultaneously reduce Ru3+ to metallic nanoparticles and transform Nb2CTx into Nb2O5 nanowires. The resulting Ru/Nb2O5/Nb2C catalyst exhibited exceptional H2 production (10.11 mmol g−1 h−1) under 313 nm illumination with 41.25% apparent quantum efficiency.

Despite these advances, the synergistic mechanisms between Nb2CTx-based materials and organic–inorganic hybrid perovskites (e.g., MAPbI3) remain underexplored. While the high electrical conductivity of Nb2CTx facilitates photogenerated electron extraction from MAPbI3, and its metallic sites may lower the energy barrier for HI decomposition, systematic studies on Nb2CTx/MAPbI3 composites for photocatalytic HI splitting are notably absent. Crucially, the synergistic optimization of interfacial stability and charge transfer kinetics in strongly acidic media represents an uncharted frontier. Addressing this knowledge gap is essential for designing perovskite-based photocatalysts with simultaneously high activity and durability.

Conventional type II heterojunctions often sacrifice redox capability for improved charge separation. To overcome this trade-off and enhance stability, we rationally selected Nb2CTx MXenes for interfacial modification of MAPbI3, driven by compelling theoretical foundations: 1) metallic conductivity and interfacial engineering: Nb2CTx exceptional metallic conductivity provides an ultra-efficient pathway for rapid electron extraction and transport from MAPbI3, minimizing recombination losses. Furthermore, it offers active proton reduction sites. The proposed in situ coupling ensures an intimate heterointerface, critical for efficient charge transfer and predicted to generate a strong built-in electric field (BIEF). This BIEF acts as an additional driving force, accelerating spatial carrier separation. 2) Z-Scheme mechanism: theoretical band alignment predicts the formation of a direct Z-scheme heterojunction at the MAPbI3/Nb2CTx interface. In this mechanism, photogenerated electrons from MAPbI3 recombine with holes in Nb2CTx (facilitated by its metallic states). This suppresses bulk recombination within MAPbI3 while preserving its highly oxidative holes and leaving highly reductive electrons on Nb2CTx for efficient proton reduction, thus maintaining strong inherent redox power. Therefore, the MAPbI3/Nb2CTx composite is theoretically designed for synergistic enhancement: the Z-scheme preserves redox capability and suppresses recombination, while Nb2CTx conductivity and interfacial engineering ensure efficient charge extraction/utilization and stability, resolving the fundamental limitations of type II systems.

This study successfully constructed a Z-scheme heterojunction MAPbI3/Nb2CTx MXenes composite photocatalytic system via an in situ coupling strategy, aiming to address the critical challenges of high carrier recombination rates and insufficient redox capability in MAPbI3 perovskite materials. Structural characterization confirmed that small-sized Nb2CTx MXenes nanosheets are uniformly anchored on the MAPbI3 surface through chemical bonding interactions, forming a heterointerface with directional charge transfer channels. Density functional theory (DFT) calculations revealed that the composite system exhibited an increased density of electronic states near the Fermi level, significantly improving the activity of the electrons participating in surface reduction reactions and thereby facilitating photocatalytic hydrogen evolution. Band structure analysis demonstrated that photogenerated electrons preferentially accumulate in the conduction band of Nb2CTx MXenes with higher reduction potentials, whereas photogenerated holes accumulate in the valence band of MAPbI3 with superior oxidation potential. This directional charge transfer pathway not only suppresses electron–hole recombination but also preserves the high reduction capability of Nb2CTx MXenes and the strong oxidation capacity of the MAPbI3 perovskite. Furthermore, the high electrical conductivity and abundant active sites on the Nb2CTx MXenes provide reaction sites for H2 evolution and accelerate interfacial charge transfer, while the strong built-in electric field enhances the carrier separation efficiency. The synergistic effects endow the composite catalyst with exceptional photocatalytic hydrogen evolution performance (12[thin space (1/6-em)]046.77 μmol h−1 g−1), retaining approximately 85% of its initial catalytic activity during consecutive cycling tests. This work establishes a novel paradigm for developing efficient and stable MAPbI3 perovskite-based composite photocatalytic systems through interface engineering and heterojunction structure design.

2. Experimental

2.1 Materials

The raw materials, including lead oxide (PbO, 99%), methylammonium iodide (MAI, CH3NH3I, 99%), hydroiodic acid (HI, 47 wt% in H2O), hydrofluoric acid (HF, 48 wt% in H2O), MAX phase compound (Nb2AlC, 400 mesh), and hypophosphorous acid (H3PO2, 50 wt% in H2O), were obtained from Sigma-Aldrich and Jilin Yiyi Technology Co., Ltd. and were used as received without further purification. Ethanol was purchased from Shaanxi China Pharmaceutical Group Co., Ltd. All the chemicals were of analytical grade and were used without additional refinement or purification. Deionized water was utilized throughout the experimental procedures.

2.2 Materials synthesis

2.2.1 Synthesis of Nb2CTx MXenes nanosheets. Two-dimensional layered Nb2CTx MXenes nanosheets were synthesized through selective etching of the Al layers from Nb2AlC in hydrofluoric acid (HF) followed by delamination. Typically, 2 g of Nb2AlC powder was gradually added to 60 mL of 40% HF solution under continuous magnetic stirring at 50 °C for 90 hours to achieve complete Al layer removal. The etched suspension was then subjected to centrifugation at 4000 rpm for 5 minutes to separate unexfoliated MXenes particles, with the supernatant collected for subsequent processing. The obtained colloidal suspension underwent repeated washing cycles with deionized water followed by centrifugation until neutralization (pH ≈ 7). The resulting sediment was thoroughly degassed with argon gas for 1 hour and subsequently exfoliated through ultrasonication at 400 W for 2 hours. The final purification involved centrifugation-assisted collection of the black precipitate, followed by ethanol washing and vacuum oven drying at 80 °C for 24 hours. This optimized procedure yielded high-quality, oxygen-terminated two-dimensional Nb2CTx MXenes nanosheets with well-defined layered structures.
2.2.2 Synthesis of MAPbI3/Nb2CTx MXenes composite photocatalytic materials by an in situ coupling method. Initially, 2 mL of hydroiodic acid (HI), 0.5 mL of H3PO2, and 0.92 g (4.12 mmol) of PbO were heated at 150 °C with magnetic stirring until complete dissolution of the yellow lead oxide powder was observed, resulting in the formation of a precursor solution. Following complete dissolution, preprepared Nb2CTx MXenes nanosheets with different mass ratios were introduced into the mixture under continuous stirring and heating for 15 minutes. Subsequently, 0.32 g (2 mmol) of MAI was added and dissolved through stirring and heating for an additional 20 minutes. Upon completion of these procedures, the reaction system was rapidly cooled to −2 °C in a Petri dish-assisted water bath. During this process in the HI solution, Nb2CTx MXenes nanosheets were anchored in situ onto the crystallized MAPbI3 surfaces, forming the MAPbI3/Nb2CTx MXenes composite photocatalytic system. Finally, the obtained MAPbI3/Nb2CTx MXenes composites were sequentially washed with ethanol and diethyl ether through three cycles of centrifugation (5000 rpm, 5 min), followed by vacuum drying at 50 °C for 24 h.

2.3 Characterizations

2.3.1 Characterization of Nb2CTx MXenes materials. Powder X-ray diffraction (XRD) measurements were performed on a Bruker D8 Advance diffractometer using Cu Kα radiation (λ = 0.15406 nm) with a 2θ range of 5–90° at a scanning rate of 10° min−1. The chemical states were investigated via X-ray photoelectron spectroscopy (XPS) recorded on a Thermo Scientific K-Alpha surface analysis system equipped with an Al Kα X-ray source. Morphological analysis of the etched material was conducted by scanning electron microscopy (SEM) images obtained via a GeminiSEM 300 microscope and transmission electron microscopy (TEM) images acquired via a JEOL JEM-F200 microscope.
2.3.2 Characterization of the MAPbI3/Nb2CTx MXenes composite photocatalytic material. The phase structure of the composite photocatalyst was analyzed via powder X-ray diffraction (XRD) on a Bruker D8 Advance diffractometer with Cu Kα radiation (λ = 0.15406 nm), covering a 2θ range of 5–90° at a scanning rate of 10° min−1. Morphological features were examined via SEM images from a TESCAN MIRA4 microscope and TEM images from a JEOL JEM-F200 microscope. The elemental composition was characterized via energy-dispersive spectroscopy (EDS) on an OXFORD XPLORE 30 microscope. The heterojunction interface between the Nb2CTx MXenes and MAPbI3 was observed through high-resolution transmission electron microscopy (HRTEM) images obtained via a JEOL JEM-F200 microscope. Chemical states were further studied by XPS on a Thermo Scientific K-Alpha system. UV-vis absorption spectra were recorded at room temperature using a Shimadzu UV-3600i Plus spectrophotometer with BaSO4 as the reference background, spanning a wavelength range of 200–800 nm. Ultraviolet photoelectron spectroscopy (UPS) measurements of different crystals were performed using a Thermo Fisher Nexsa system.

2.4 Photocatalytic H2 evolution

Unless otherwise specified, all photocatalytic hydrogen evolution experiments were conducted via a photochemical reactor (Model MC-SPH 2O-A, Merry Change) coupled with an online gas chromatograph (GC 9790 II, FULI INSTRUMENTS) for real-time detection. The reactor system comprises an outer housing and a separate external water-cooled circulation system. The housing contains an 80 mL Pyrex glass reaction chamber, while the external cooling system is equipped with a quartz window. The experimental parameters were standardized as follows: illumination area of 28.26 cm2, light intensity of 100 mW cm−2 provided by a light source positioned 12 cm above the reaction solution surface, operating current of 10 mA, and irradiation duration of 5 h. Hydrogen production was quantified through GC analysis via a pre-established standard calibration curve. The typical experimental procedure consisted of four steps: 1) 200 mg of vacuum-dried MAPbI3/Nb2CTx MXenes composite was mixed with 20 mL of saturated HI/H3PO2 aqueous solution in the reaction chamber; 2) the reaction system was cooled to 6 °C and evacuated under vacuum for 10 min; 3) the absence of residual H2 was confirmed by an online GC prior to illumination; and 4) the photocatalytic HI splitting reaction was initiated upon light irradiation, while H2 generation was continuously monitored through GC analysis.

2.5 Calculation method

The computational simulations in this study were conducted via density functional theory (DFT) implemented in the Vienna Ab initio Simulation Package (VASP),56 which uses projector augmented wave (PAW) pseudopotentials.57 Electron exchange–correlation interactions were described through the Perdew–Burke–Ernzerhof (PBE)56 generalized gradient approximation enhanced with semiempirical dispersion correction (DFT-D).58 Spin-polarized calculations were carried out with a plane-wave energy cutoff of 520 eV to ensure basis set completeness. To enhance the structural parameters and refine the unit cell dimensions, a Gamma-centered59 9 × 9 × 1 k-point mesh was employed during numerical integration across the Brillouin zone. The electronic self-consistent field iteration process maintained a rigorous convergence criterion of 10–5 eV per atom for total energy differences. All optimized configurations satisfied atomic force thresholds below 0.02 eV Å−1 while maintaining residual stress components within computational tolerance limits.

3. Results and discussion

3.1 Characterization of the structures and morphology of the Nb2CTx MXenes

The Nb2CTx MXenes nanosheets were synthesized by etching the Nb2AlC MAX phase precursor with HF, as illustrated in Fig. 1(a). As evidenced by the XRD patterns of the materials before and after etching (Fig. 1(b)), the (002) diffraction peak of the Nb2CTx MXenes shifted to a lower angle after HF treatment, indicating an expansion of the interlayer spacing due to the removal of Al atoms and the synergistic effects of surface functional groups (–OH, –F, and –O) and intercalated water molecules, which led to structural loosening of the originally closely packed Nb2AlC layers.60 Concurrently, the exposed Nb2CTx MXenes surface underwent functionalization with these moieties. XPS analysis confirmed the successful synthesis of the Nb2CTx MXenes (Fig. 1(c)). The total XPS spectrum is shown in Fig. S1. The near-complete disappearance of the Al 2p signal at ∼74 eV verified the effective removal of the Al layer. In the Nb 3d spectrum, the intensity of the original Nb–C bond (206.9 eV, 3d5/2) decreased significantly, whereas new peaks emerged at 206.9 eV and 210.0 eV, corresponding to Nb–F and Nb–O bonds, respectively, indicating surface fluorination/oxidation. The O 1s spectrum revealed contributions from Nb–O (530.0 eV) and C[double bond, length as m-dash]O (533.1 eV). These results collectively verify the successful synthesis of Nb2CTx MXenes with abundant surface –O and –F functionalities. The SEM and TEM images of the precursor MAX phase and MXenes phase before and after HF acid etching are shown in Fig. S2(a and b). The SEM image combined with the corresponding EDS map presented in Fig. 1(d) reveals that the Nb2CTx MXenes material has been uniformly etched into a characteristic accordion-like morphology. Furthermore, the TEM image in Fig. 1(e) further confirms the formation of a layered structure. The structural stability of the Nb2CTx MXenes in HI aqueous solution was evaluated via X-ray diffraction (XRD). Minimal changes in the XRD patterns before and after HI immersion (Fig. S3) demonstrated the excellent chemical stability of the Nb2CTx MXenes under HI treatment.
image file: d5lf00214a-f1.tif
Fig. 1 (a) Etching process schematic of the Nb2CTx MXenes phase material obtained via HF acid etching of the Nb2AlC MAX phase precursor. (b) XRD images of the Nb2AlC MAX phase and Nb2CTx MXenes phase before and after etching. (c) XPS images, (d) SEM images, EDS images and (e) TEM images of the etched Nb2CTx MXenes phase.

3.2 Characterization of the structure and morphology of MAPbI3/Nb2CTx MXenes

The heterojunction photocatalysts of MAPbI3 and Nb2CTx MXenes nanosheets (MAPbI3/Nb2CTx MXenes) were fabricated via an in situ coupling method. Specifically, various quantities (20 mg, 40 mg, 80 mg, 100 mg, and 120 mg) of synthesized Nb2CTx MXenes were introduced into the precursor solution for MAPbI3 perovskite synthesis. The resulting composites were designated MAPbI3/20MX, MAPbI3/40MX, MAPbI3/80MX, MAPbI3/100MX and MAPbI3/120MX, respectively. The as-prepared MAPbI3/Nb2CTx MXenes composites exhibited a structural configuration in which multiple small-sized Nb2CTx MXenes nanosheets were anchored onto larger MAPbI3 substrates (Fig. 2(a)). This architecture endows the composites with enhanced charge separation efficiency and superior photocatalytic performance for HI decomposition.
image file: d5lf00214a-f2.tif
Fig. 2 (a) Schematic illustration of the structural configuration and redox processes of MAPbI3/Nb2CTx MXenes. (b) SEM images of the MAPbI3/Nb2CTx MXenes composite photocatalyst with 40 mg of Nb2CTx MXenes added. (c) SEM images and corresponding EDX maps of MAPbI3/Nb2CTx MXenes. (d) XRD images of pure MAPbI3, pure Nb2CTx MXenes and MAPbI3/Nb2CTx MXenes composites with 40 mg of Nb2CTx MXenes added. (e) TEM/HRTEM images of MAPbI3/Nb2CTx MXenes after compositing.

SEM characterization revealed the morphological features of the composite materials. Smooth rhombic dodecahedral MAPbI3 crystals were densely decorated with multilayered Nb2CTx MXenes nanosheets, forming well-integrated MAPbI3/Nb2CTx MXenes composites (Fig. 2(b)). SEM images of the MAPbI3/Nb2CTx MXenes composites modified with different masses of Nb2CTx MXenes are shown in Fig. S4. EDS elemental mapping confirmed the homogeneous distribution of Pb, I, Nb, and C throughout the composite structure, verifying the successful incorporation of Nb2CTx MXenes into the MAPbI3 matrix (Fig. 2(c)).

XRD was employed to investigate the crystalline structures of the MAPbI3 and Nb2CTx MXenes and their composites (Fig. 2(d)). The XRD pattern of pristine MAPbI3 exhibits characteristic peaks at 14.1, 28.4, 31.8, 40.4, and 43.0 nm, which are indexed to the (110), (220), (310), (400), and (314) crystallographic planes of the tetragonal perovskite structure, respectively.10 The XRD patterns of the MAPbI3/Nb2CTx MXenes composites modified with different masses of Nb2CTx MXenes and the MAPbI3/Nb2CTx MXenes composites modified with different sizes are shown in Fig. S5 and S6. Notably, the MAPbI3/Nb2CTx MXenes composite material retains the characteristic diffraction peaks corresponding to both Nb2CTx MXenes and pure MAPbI3 components. Furthermore, XRD analysis confirmed that the MAPbI3/Nb2CTx MXenes hybrid maintains a tetragonal perovskite phase identical to that of pristine MAPbI3, demonstrating the structural integrity of the perovskite framework during composite formation.

TEM observations further confirmed the successful coupling between the Nb2CTx MXenes nanosheets and MAPbI3 (Fig. 2(e)). HRTEM imaging revealed interplanar spacings of 0.576 nm, 0.362 nm, and 0.980 nm, which correspond to the (110) plane of MAPbI3 and the (202)/(002) planes of Nb2CTx MXenes, respectively. These structural characterizations collectively demonstrate the successful fabrication of MAPbI3/Nb2CTx MXenes heterojunction photocatalysts with well-defined interfacial contacts.

3.3 Photocatalytic HER activities of MAPbI3/Nb2CTx MXenes

Fig. 3(a) shows the photocatalytic hydrogen production rates and corresponding hydrogen production rates of the MAPbI3/Nb2CTx MXenes composites with different masses of Nb2CTx MXenes. Fig. 3(b) shows the photocatalytic hydrogen production and corresponding hydrogen production rates of the MAPbI3/Nb2CTx MXenes composites with different sizes of Nb2CTx MXenes. The HER performance of the MAPbI3/Nb2CTx MXenes photocatalysts in a 20 mL aqueous HI solution was evaluated under visible light irradiation (λ ≥ 420 nm) using H3PO2 as a sacrificial electron donor. As shown in Fig. 3a, although pure MAPbI3 exhibits photocatalytic activity, its HER rate remains as low as 35 μmol h−1 g−1 because of the rapid recombination of photogenerated carriers. After anchoring HF-etched Nb2CTx MXenes nanosheets (predominantly 2 μm in lateral size) onto MAPbI3, the HER activity for HI splitting was significantly enhanced. Compared with pristine MAPbI3, MXenes with varying mass ratios demonstrated substantially higher H2 evolution rates. Specifically, MAPbI3–20MX, MAPbI3–40MX, MAPbI3–80MX, MAPbI3-100MX and MAPbI3-120MX presented H2 production rates of 4251.97, 7294.53, 2633.26, 729.53, and 1078.75 μmol h−1 g−1, respectively, corresponding to 121-, 208-, 75-, 20-, and 31-fold enhancements over those of pure MAPbI3. This confirms the critical role of Nb2CTx MXenes loading in boosting HER performance for HI splitting. Further optimization was achieved by reducing the lateral size of the optimal 40 mg of Nb2CTx MXenes (initially 2 μm) to ∼1 μm through thermal agitation and ultrasonication. The resulting MAPbI3–40MX-1 μm photocatalyst delivered a remarkable HER rate of 12046.77 μmol h−1 g−1 (200.78 μmol h−1), representing 344-fold and 1.65-fold improvements over those of pristine MAPbI3 (ref. 61) (35 μmol h−1 g−1) and MAPbI3–40MX (7294.53 μmol h−1 g−1), respectively (Fig. S7). This enhancement arises from the exposure of unsaturated coordination sites at active edges and improved electron transfer properties in smaller MXenes nanosheets.62,63 However, excessive size reduction to ∼500 nm (MAPbI3–40MX-500 nm) drastically decreased the HER performance due to synergistic effects, including quantum confinement, increased surface defects, and nanosheet aggregation.
image file: d5lf00214a-f3.tif
Fig. 3 (a) Photocatalytic hydrogen production rates and corresponding hydrogen production rates of MAPbI3/Nb2CTx MXenes composites with different masses of Nb2CTx MXenes (under visible light, the photocatalytic hydrogen production of the sample within 5 h). (b) Photocatalytic hydrogen production rates and corresponding hydrogen production rates of MAPbI3/Nb2CTx MXenes composites with different sizes of Nb2CTx MXenes (under visible light, the photocatalytic hydrogen production of the sample within 5 h). (c) Cycling tests of the H2 production performance of MAPbI3/40MX.

To assess practical applicability, five consecutive recycling tests (5 h per cycle) were conducted under identical conditions (Fig. 3(c)). MAPbI3–40MX maintained stable HER activity throughout the cycles, demonstrating excellent photostability. The minor activity decline in later cycles likely originated from catalyst deposition on magnetic stir bars. Postreaction characterization revealed nearly identical XRD patterns (Fig. S8(a)) and SEM images (Fig. S8(b)) before and after cycling, confirming the structural integrity and compositional stability of MAPbI3–40MX during photocatalytic operation. These results highlight the exceptional photocatalytic durability and structural robustness of the MAPbI3/Nb2CTx MXenes composite system.

Compared with the current literature data, as shown in Fig. S9, the pink bars represent the HER yields of MXenes-based composites, whereas the orange bars represent those of perovskite MAPbI3-based composites. The figure shows that the HI-splitting HER efficiency of the MAPbI3/Nb2CTx MXenes composite developed in this work is in the upper-middle range, indicating its promising potential as an excellent photocatalytic hydrogen production material.

Further investigation into the photocatalytic performance of MAPbI3 composites with other MXenes (Ti-, Mo-, W-, and Ta-based MXenes) under identical conditions (1 μm particle size) is presented in Fig. S10. Notably, the Nb2CTx–MAPbI3 composite demonstrates the highest HER activity among all tested MXenes-based composites. To validate the role of the Z-scheme heterojunction in enhancing performance, control experiments were conducted using pure Nb2CTx MXenes as the photocatalyst (Fig. S11). Strikingly, the HER efficiency of the Nb2CTx–MAPbI3 composite exceeds that of pure Nb2CTx MXenes by over ninefold. This significant enhancement provides strong evidence for the formation and efficacy of the Z-scheme heterojunction structure in facilitating charge separation and boosting photocatalytic hydrogen evolution.

The solar light harvesting capabilities of the Nb2CTx MXenes nanosheets, MAPbI3, and MAPbI3/Nb2CTx MXenes composites were evaluated through UV-vis diffuse reflectance spectroscopy (Fig. 4(a) and S12). The Nb2CTx MXenes (black curve) exhibit a broad absorption spectrum without distinct absorption edges, demonstrating full-spectrum absorption characteristics consistent with their metal-like properties. MAPbI3 displays visible light absorption due to its semiconductor nature. Consequently, the MAPbI3/Nb2CTx MXenes composite shows strong absorption between 200 and 520 nm, enabling efficient harvesting of visible light for photocatalytic HI splitting toward hydrogen generation.


image file: d5lf00214a-f4.tif
Fig. 4 (a) UV-vis spectra and band gap energies of composite Nb2CTx MXenes and MAPbI3 at different ratios. (b) SPV signals of composite Nb2CTx MXenes and MAPbI3 at different ratios. (c) UPS spectra of composite Nb2CTx MXenes and MAPbI3 at different ratios. (d) Schematic band diagram of MAPbI3/Nb2CTx MXenes for photocatalytic HI splitting reactions.

The carrier separation efficiency was investigated via surface photovoltage (SPV) spectroscopy, which directly detects the photogenerated voltage under illumination, with the SPV signal intensity reflecting the charge separation capability. As shown in Fig. 4(b), the SPV signals of MAPbI3/40MX (35.37 ± 2.6 mV) and MAPbI3/40MX-1 μm (47.43 ± 3.1 mV) significantly surpassed those of MAPbI3/20MX (29.27 ± 1.8 mV), MAPbI3/80MX (27.77 ± 3.0 mV), MAPbI3/100MX (12.56 ± 3.0 mV), MAPbI3/120MX (10.80 ± 1.4 mV), and MAPbI3/40MX-500 nm (6.8 ± 0.8 mV). This clearly demonstrates enhanced charge separation in MAPbI3 modified with 40 mg of Nb2CTx MXenes. Notably, MAPbI3/40MX-1 μm has the highest SPV response, suggesting an optimal charge distribution that contributes to its superior hydrogen evolution performance.

Fig. 4(c) shows the ultraviolet photoelectron spectroscopy (UPS) spectra of the crystal, whereas Fig. 4d shows the energy level diagram of the MAPbI3/Nb2CTx MXenes-based HI-splitting hydrogen production system constructed from the UPS data. The inset in the upper right corner of Fig. 4(d) schematically depicts the electron–hole separation in the Z-scheme heterojunction photocatalyst under light irradiation.64 As revealed in Fig. 4(d), the Z-scheme heterojunction formed between the MAPbI3 and Nb2CTx MXenes enables more efficient photoinduced charge separation and enhanced reduction capacity than conventional type II heterojunction photocatalysts do. In this Z-scheme configuration, the conduction band (CB) and valence band (VB) of the perovskite material exhibit lower energy levels than those of typical semiconductor materials. The experimental results demonstrate that the MAPbI3/40MX-1 μm heterojunction (with the incorporation of 1 μm-sized Nb2CTx MXenes) achieves optimized band alignment, with a valence band maximum (VBM) of −4.61 eV and a conduction band minimum (CBM) of −2.49 eV. Notably, the more negative VB position of this heterojunction compared with that of pristine MAPbI3 (−4.47 eV vs. vacuum level (VAC) for the H+/H2 redox potential) meets the thermodynamic requirements for the photocatalytic HER. In the Z-scheme heterojunction, photogenerated electrons spontaneously transfer from the CB of MAPbI3 to the VB of Nb2CTx MXenes for hole recombination. This directional charge migration results in electron accumulation on the higher-reduction-potential CB of the MXenes and hole retention on the stronger-oxidation-capacity VB of MAPbI3. Such charge separation dynamics not only suppress electron–hole recombination but also preserve the superior reduction capability of MXenes and the robust oxidation capacity of perovskite, thereby synergistically driving H+ reduction and HI oxidation reactions. Furthermore, the high conductivity and abundant surface active sites of the Nb2CTx MXenes provide efficient reaction sites for H2 evolution while accelerating interfacial charge transfer. The strengthened built-in electric field significantly enhances the carrier separation efficiency. Concurrently, the photogenerated holes in the VB of MAPbI3 oxidize I to I3, which is subsequently consumed by H3PO2 during photocatalytic processes, maintaining solution stability. Stability tests confirmed the structural integrity of MAPbI3/40MX-1 μm after 12 hours of continuous photocatalytic operation, with SEM images and XPS analysis showing no significant perovskite degradation or MXenes oxidation (Fig. S13), demonstrating the synergistic stabilization effect of the Z-scheme heterojunction.

To further elucidate the photogenerated carrier transfer pathway in the MAPbI3/Nb2CTx MXenes composite, in situ XPS characterization was performed under illumination. As shown in Fig. 5(a), upon light irradiation, the binding energies of the Pb 4f peaks in the MAPbI3/Nb2CTx MXenes composite shifted from 143.02, 138.14, and 134.92 eV to 143.33, 138.44, and 135.38 eV, respectively. Similarly, the I 3d peak shifted from 618.85 and 630.36 eV to 619.21 and 630.69 eV. This positive shift in binding energies suggests the accumulation of holes within the MAPbI3 component following photoexcitation.


image file: d5lf00214a-f5.tif
Fig. 5 (a) High-resolution XPS spectra of Pb 4f, I 3d, Nb 3d and C 1s of MAPbI3/Nb2CTx MXenes before and after illumination. Illustration of the band structures between Nb2CTx MXenes and MAPbI3. (b) before contact, (c) in the dark, and (d) under light irradiation. The Ef bending at the interface was considered in (c) and (d).

Conversely, the binding energies of the Nb 3d peaks decreased from 204.65 and 207.28 eV to 204.44 and 207.11 eV, whereas the C 1s peaks shifted from 286.44 and 288.75 eV to 286.29 and 288.35 eV. This negative shift is attributed to the accumulation of negatively charged photogenerated electrons within the Nb2CTx MXenes component.

Considering the band alignment of the MAPbI3 and Nb2CTx MXenes, these observations collectively indicate that, upon illumination, photogenerated electrons from MAPbI3 transfer to the Nb2CTx MXenes. This electron transfer serves two critical functions: suppressing electron–hole recombination within the Nb2CTx MXenes by annihilating holes. Facilitates spatial charge separation, leading to the accumulation of a high density of photogenerated electrons in the Nb2CTx MXenes and a corresponding accumulation of photogenerated holes in the MAPbI3.

Consequently, the observed XPS shifts and the proposed charge transfer mechanism provide compelling evidence for a direct Z-scheme (or Z-type) heterojunction within the MAPbI3/Nb2CTx MXenes composite, where the photogenerated electrons follow a Z-scheme transfer pathway.

Following the formation of the MAPbI3/Nb2CTx MXenes heterojunction, the lower work function and higher Fermi level of MAPbI3 MXenes than of Nb2CTx MXenes (Fig. 5(b)) create thermodynamically favorable conditions for spontaneous electron transfer from MAPbI3 to Nb2CTx MXenes. This electron migration persists until the Fermi levels of the two components equilibrate. Consequently, electron depletion and accumulation layers are established at the interface under equilibrium conditions. This results in a positively charged surface on MAPbI3 and a negatively charged surface on Nb2CTx MXenes, inducing the formation of a robust interfacial electric field (IEF) and band bending at the MAPbI3/Nb2CTx MXenes interface (Fig. 5(c)). This configuration provides a potent driving force for interfacial charge transfer.

Under light illumination, the IEF promotes the recombination of photogenerated electrons in the CBM of Nb2CTx MXenes with photogenerated holes in the VBM of MAPbI3 via an S-shaped electron transfer pathway (Fig. 5(d)). Consequently, photogenerated electrons are preserved in the CBM of MAPbI3, whereas photogenerated holes accumulate in the VBM of Nb2CTx MXenes. Thermodynamically, the estimated redox potentials of the MAPbI3/Nb2CTx MXenes heterojunction exhibit overlapping regions with the theoretical potentials required to simultaneously drive reduction reactions on the Nb2CTx MXenes side and oxidation reactions on the MAPbI3 side under visible light irradiation.

3.4 Photocatalytic mechanism

The interfacial charge transfer characteristics between the MAPbI3 and Nb2CTx MXenes were systematically investigated via DFT calculations. Fig. S14 presents the optimized configurations of pristine MAPbI3 and Nb2CTx MXenes and their heterostructures (denoted MAPbI3/Nb2CTx MXenes). DFT-derived electronic band structures (Fig. S15) and density of states (DOS) profiles (Fig. 6(a)) reveal distinct semiconductor characteristics in MAPbI3 and quasi-metallic behavior in Nb2CTx MXenes. To elucidate the interfacial charge redistribution, differential charge density analysis was performed at the MAPbI3/Nb2CTx MXenes interface (Fig. 6(b)). The planar-averaged charge density difference along the Z-axis demonstrates significant electron accumulation (yellow regions) localized on the Nb2CTx MXenes component, whereas electron depletion (cyan regions) predominantly occurs in the MAPbI3 region, indicative of directional electron transfer from MAPbI3 to Nb2CTx MXenes. This charge transfer mechanism is further corroborated by work function calculations (Fig. 6(c)), which yield values of 4.571 eV for MAPbI3 and 5.77 eV for the Nb2CTx MXenes. The higher work function of the Nb2CTx MXenes establishes an electron-driving force at the interface, which is consistent with the observed differential charge distribution and confirms the formation of a Z-scheme heterojunction architecture. Furthermore, hydrogen adsorption Gibbs free energy (ΔGH*) calculations (Fig. 6(d)) demonstrated remarkable catalytic enhancement in the composite system. The optimized MAPbI3/Nb2CTx MXenes heterostructure exhibited near-zero |ΔGH*| values, suggesting a significantly reduced overpotential and improved HER activity during HI photocatalytic decomposition. This synergistic effect arises from the Z-scheme charge transfer pathway, which effectively suppresses carrier recombination while maintaining strong redox potentials.65
image file: d5lf00214a-f6.tif
Fig. 6 (a) The densities of states of the MAPbI3 and Nb2CTx MXenes are calculated. (b) The differential charge density distribution at the interface of the MAPbI3/Nb2CTx MXenes complex was simulated. (c) Work function calculation of pure MAPbI3, pure Nb2CTx MXenes and MAPbI3-MX. (d) Gibbs free energy calculation of H atom adsorption on pure MAPbI3, pure Nb2CTx MXenes and MAPbI3-MX.

Conclusions

This study successfully constructed a Z-scheme heterojunction-based MAPbI3/Nb2CTx MXene composite photocatalyst for hydrogen production via HI splitting. An in situ coupling strategy enabled uniform anchoring of Nb2CTx nanosheets onto MAPbI3 with interfacial chemical bonding, establishing intimate contact and directional charge-transfer channels. The composite achieved exceptional photocatalytic hydrogen evolution (12[thin space (1/6-em)]046.77 μmol h−1 g−1), representing a 344-fold enhancement over pristine MAPbI3, while retaining 85% activity after cycling-overcoming the efficiency–stability trade-off typical of perovskite photocatalysts. Theoretical analyses reveal the synergistic mechanism: a robust interfacial built-in electric field drives Z-scheme charge migration, facilitating the participation of highly reductive electrons from the Nb2CTx conduction band and oxidative holes from the MAPbI3 valence band in surface reactions. The metallic conductivity of MXenes accelerates charge kinetics, while their abundant active sites optimize proton reduction. This work demonstrates Nb2CTx as an efficient noble-metal-free cocatalyst and establishes a structure-interface-property paradigm for designing stable, high-performance photocatalytic systems.

Author contributions

Ruiyan Sun: conceptualization, data curation, formal analysis, and writing – original draft. Lili Gao: funding acquisition, investigation, methodology, project administration, resources, supervision, and writing – review & editing. Deng Li: data curation, formal analysis, investigation, supervision, and writing – review & editing. Hua Wang: investigation, supervision, and writing – review & editing. Fan Yang: investigation, methodology, supervision, and writing –review & editing. Jin Wang: investigation, methodology, project administration, and writing – review & editing. Ke Hao: data curation, formal analysis, and writing – original draft. Haijiao Xie: data curation, and formal analysis. Ping Hu: funding acquisition, investigation, methodology, project administration, resources, supervision, and writing – review & editing.

Conflicts of interest

There are no conflicts to declare.

Data availability

Supplementary information: Figures: XPS survey spectrum of Nb2CTx MXenes after etching; SEM patterns and TEM patterns before and after HF-etched; XRD patterns before and after HI immersion; SEM images of Nb2CTx MXenes modified MAPbI3/Nb2CTx MXenes composites with different mass; XRD patterns of MAPbI3/Nb2CTx MXenes composites modified with different mass of Nb2CTx MXenes; XRD patterns of Nb2CTx MXenes modified MAPbI3/Nb2CTx MXenes composites with different sizes; the HER rates of the original MAPbI3, MAPbI3/40MX and MAPbI3/40MX-1 μm photocatalysts were compared; XRD patterns of MAPbI3/40MX before and after cycling; compared with other MXenes-based (pink) and MAPbI3-based (orange) photocatalytic hydrogen evolution systems; the photocatalytic hydrogen evolution results of MAPbI3 doped with different particle sizes of 1 μm MXenes (Ti-, Mo-, W-, Ta-based); the photocatalytic hydrogen evolution performance of pure Nb2CTx MXenes and Nb2CTx–MAPbI3 (40MX-1 μm) after compounding with MAPbI3 was compared; UV-vis diffuse reflectance spectra of Nb2CTx MXenes nanosheets and MAPbI3/Nb2CTx MXenes composites; SEM image and XPS analysis of MAPbI3/40MX-1 μm before and after 12 hours of continuous photocatalytic hydrogen production; the configuration diagrams of pure MAPbI3, pure Nb2CTx MXenes and MAPbI3-MX after compounding; the band structure of MAPbI3 and Nb2CTx MXenes are calculated. See DOI: https://doi.org/10.1039/D5LF00214A.

All data supporting the findings of this study are fully presented within the manuscript text and figures. No additional external datasets were generated or analyzed, except for the comparative data used in Fig. S9. The data for Fig. S9 were derived from the study by Li et al. 2022 (DOI: https://doi.org/10.1016/j.jpowsour.2022.231006) and Nemamcha et al. 2024 (DOI: https://doi.org/10.1016/j.scitotenv.2024.172816).

Acknowledgements

This work was supported by the “Young Talent Support Project” of the Shaanxi Association for Science and Technology (20240438), the Key R&D Plan of Shaanxi Province (2025QCY-KXJ-008), the Natural Science Foundation of Shaanxi Province (2025JC-YBQN-483), the Scientific Research Program of Youth Innovation Team of Shaanxi (23JC045), the Scientific Research Program of Youth Innovation Team of Shaanxi (24JP094) and the State Key Laboratory for Mechanical Behavior of Materials.

References

  1. S. K. Lakhera, A. Rajan, T. P. Rugma and N. Bernaurdshaw, Renewable Sustainable Energy Rev., 2021, 152, 111694 CrossRef.
  2. T. Hisatomi and K. Domen, Nat. Catal., 2019, 2, 387–399 CrossRef.
  3. J. H. Kim, D. Hansora, P. Sharma, J.-W. Jang and J. S. Lee, Chem. Soc. Rev., 2019, 48, 1908–1971 RSC.
  4. W. Zhang, J. Xiong, S. Li and W. Li, Mol. Catal., 2025, 570, 114701 Search PubMed.
  5. G. Cheng, X. Liu and J. Xiong, Chem. Eng. J., 2024, 501, 157491 Search PubMed.
  6. P. Qiu, J. Xiong, M. Lu, L. Liu, W. Li, Z. Wen, W. Li, R. Chen and G. Cheng, J. Colloid Interface Sci., 2022, 622, 924–937 CrossRef PubMed.
  7. Y. Wei, G. Cheng, J. Xiong, J. Zhu, Y. Gan, M. Zhang, Z. Li and S. Dou, J. Energy Chem., 2019, 32, 45–56 CrossRef.
  8. C. Wang, J. Xiong, Z. Wen and G. Cheng, Ind. Eng. Chem. Res., 2023, 62, 11402–11413 CrossRef.
  9. P. Zhou, H. Chen, Y. Chao, Q. Zhang, W. Zhang, F. Lv, L. Gu, Q. Zhao, N. Wang, J. Wang and S. Guo, Nat. Commun., 2021, 12, 4412 CrossRef PubMed.
  10. Z. Zhao, J. Wu, Y.-Z. Zheng, N. Li, X. Li and X. Tao, ACS Catal., 2019, 9, 8144–8152 CrossRef.
  11. X. Wang, H. Wang, H. Zhang, W. Yu, X. Wang, Y. Zhao, X. Zong and C. Li, ACS Energy Lett., 2018, 3, 1159–1164 CrossRef.
  12. M. Wang, Y. Zuo, J. Wang, Y. Wang, X. Shen, B. Qiu, L. Cai, F. Zhou, S. P. Lau and Y. Chai, Adv. Energy Mater., 2019, 9, 1901801 CrossRef.
  13. S. Park, W. J. Chang, C. W. Lee, S. Park, H.-Y. Ahn and K. T. Nam, Nat. Energy, 2017, 2, 16185 CrossRef.
  14. R. Li, X. Li, J. Wu, X. Lv, Y.-Z. Zheng, Z. Zhao, X. Ding, X. Tao and J.-F. Chen, Appl. Catal., B, 2019, 259, 118075 CrossRef.
  15. R. Marschall, Adv. Funct. Mater., 2014, 24, 2421–2440 CrossRef.
  16. S. Bai, J. Jiang, Q. Zhang and Y. Xiong, Chem. Soc. Rev., 2015, 44, 2893–2939 RSC.
  17. Q. Wang and K. Domen, Chem. Rev., 2020, 120, 919–985 CrossRef.
  18. B. Li, W. Guo, X. F. Lu, Y. Hou, Z. Ding and S. Wang, Mater. Rep.: Energy, 2023, 3, 100230 Search PubMed.
  19. B. Li, W. Wang, J. Zhao, Z. Wang, B. Su, Y. Hou, Z. Ding, W.-J. Ong and S. Wang, J. Mater. Chem. A, 2021, 9, 10270–10276 RSC.
  20. J. Cai, X. Li, B. Su, B. Guo, X. Lin, W. Xing, X. F. Lu and S. Wang, J. Mater. Sci. Technol., 2025, 234, 82–89 CrossRef.
  21. X. Zhao, S. Chen, H. Yin, S. Jiang, K. Zhao, J. Kang, P. F. Liu, L. Jiang, Z. Zhu, D. Cui, P. Liu, X. Han, H. G. Yang and H. Zhao, Matter, 2020, 3, 935–949 CrossRef.
  22. A. Kojima, K. Teshima, Y. Shirai and T. Miyasaka, J. Am. Chem. Soc., 2009, 131, 6050–6051 CrossRef PubMed.
  23. S. D. Stranks, G. E. Eperon, G. Grancini, C. Menelaou, M. J. P. Alcocer, T. Leijtens, L. M. Herz, A. Petrozza and H. J. Snaith, Science, 2013, 342, 341–344 CrossRef.
  24. C. Wehrenfennig, G. E. Eperon, M. B. Johnston, H. J. Snaith and L. M. Herz, Adv. Mater., 2014, 26, 1584–1589 CrossRef PubMed.
  25. Q. Dong, Y. Fang, Y. Shao, P. Mulligan, J. Qiu, L. Cao and J. Huang, Science, 2015, 347, 967–970 CrossRef.
  26. Q. Guo, C.-Y. Wang, T. Hayat, A. Alsaedi, J.-X. Yao and Z.-A. Tan, Rare Met., 2021, 40, 2763–2777 CrossRef.
  27. J. Ran, J. Zhang, J. Yu, M. Jaroniec and S. Z. Qiao, Chem. Soc. Rev., 2014, 43, 7787–7812 RSC.
  28. Y. Wu, P. Wang, X. Zhu, Q. Zhang, Z. Wang, Y. Liu, G. Zou, Y. Dai, M.-H. Whangbo and B. Huang, Adv. Mater., 2018, 30, 1704342 CrossRef.
  29. Y. Zhang, J. Shi, X. Ding, J. Wu, Y.-Z. Zheng and X. Tao, Ind. Eng. Chem. Res., 2020, 59, 20667–20675 CrossRef.
  30. F. Wang, X. Liu, Z. Zhang and S. Min, Chem. Commun., 2020, 56, 3281–3284 RSC.
  31. Y. Liu, M. Tayyab, W. Pei, L. Zhou, J. Lei, L. Wang, Y. Liu and J. Zhang, Small, 2023, 19, 2208117 CrossRef.
  32. Y. Sun, D. Jin, Y. Sun, X. Meng, Y. Gao, Y. Dall'Agnese, G. Chen and X.-F. Wang, J. Mater. Chem. A, 2018, 6, 9124–9131 RSC.
  33. Y. Li, L. Ding, S. Yin, Z. Liang, Y. Xue, X. Wang, H. Cui and J. Tian, Nano-Micro Lett., 2020, 12, 4998 Search PubMed.
  34. J. Wang, Y. Shen, S. Liu and Y. Zhang, Appl. Catal., B, 2020, 270, 118885 CrossRef.
  35. G. Zuo, Y. Wang, W. L. Teo, A. Xie, Y. Guo, Y. Dai, W. Zhou, D. Jana, Q. Xian, W. Dong and Y. Zhao, Angew. Chem., Int. Ed., 2020, 59, 11287–11292 CrossRef PubMed.
  36. C. Chen, X. Xie, B. Anasori, A. Sarycheva, T. Makaryan, M. Zhao, P. Urbankowski, L. Miao, J. Jiang and Y. Gogotsi, Angew. Chem., Int. Ed., 2018, 57, 1846–1850 CrossRef PubMed.
  37. J. Shen, G. Liu, Y. Ji, Q. Liu, L. Cheng, K. Guan, M. Zhang, G. Liu, J. Xiong, J. Yang and W. Jin, Adv. Funct. Mater., 2018, 28, 1801511 CrossRef.
  38. E. M. S. Aldeen, A. A. Jalil, R. S. Mim, A. H. Hatta, N. I. H. Hazril, A. Chowdhury, N. S. Hassan and S. Rajendran, Environ. Res., 2023, 234, 116576 CrossRef.
  39. R. Sun, P. Hu, J. Wang, F. Yang, F. Zhu, H. Xing, J. Luo, L. Gao, K. Wang and Z. Yin, Small, 2025, 21, 2408331 CrossRef.
  40. J. Ran, G. Gao, F.-T. Li, T.-Y. Ma, A. Du and S.-Z. Qiao, Nat. Commun., 2017, 8, 13907 CrossRef.
  41. B. Wang, M. Wang, F. Liu, Q. Zhang, S. Yao, X. Liu and F. Huang, Angew. Chem., Int. Ed., 2020, 59, 1914–1918 CrossRef PubMed.
  42. H. Wang, R. Peng, Z. D. Hood, M. Naguib, S. P. Adhikari and Z. Wu, ChemSusChem, 2016, 9, 1490–1497 CrossRef PubMed.
  43. T. Ke, S. Shen, K. Rajavel, K. Yang and D. Lin, J. Hazard. Mater., 2021, 402, 124066 CrossRef.
  44. Z. Ai, K. Zhang, L. Xu, M. Huang, D. Shi, Y. Shao, J. Shen, Y. Wu and X. Hao, J. Colloid Interface Sci., 2022, 610, 13–23 CrossRef PubMed.
  45. M. Shao, Y. Shao, J. Chai, Y. Qu, M. Yang, Z. Wang, M. Yang, W. F. Ip, C. T. Kwok, X. Shi, Z. Lu, S. Wang, X. Wang and H. Pan, J. Mater. Chem. A, 2017, 5, 16748–16756 RSC.
  46. H. Pan, Sci. Rep., 2016, 6, 32531 CrossRef PubMed.
  47. G. Gao, A. P. O'Mullane and A. Du, ACS Catal., 2017, 7, 494–500 CrossRef.
  48. I. Hussain, A. Hanan, F. Bibi, O. J. Kewate, M. S. Javed and K. Zhang, Adv. Energy Mater., 2024, 14, 2401650 CrossRef.
  49. S. Yang, P. Zhang, A. S. Nia and X. Feng, Adv. Mater., 2020, 32, 1907857 CrossRef.
  50. H. Wang, Y. Wu, X. Yuan, G. Zeng, J. Zhou, X. Wang and J. W. Chew, Adv. Mater., 2018, 30, 1704561 CrossRef.
  51. G. Guan and F. Guo, Batteries, 2023, 9, 235 CrossRef.
  52. M. Tayyab, Y. Liu, Z. Liu, L. Pan, Z. Xu, W. Yue, L. Zhou, J. Lei and J. Zhang, J. Colloid Interface Sci., 2022, 628, 500–512 CrossRef.
  53. M. Tayyab, U. E. Kulsoom, Y. Liu, S. Mansoor, M. Khan, Z. Akmal, A. Mushtaq, M. Arif, U. Shamriaz, L. Zhou, J. Lei and J. Zhang, Int. J. Hydrogen Energy, 2024, 51, 1400–1413 CrossRef.
  54. T. Su, R. Peng, Z. D. Hood, M. Naguib, I. N. Ivanov, J. K. Keum, Z. Qin, Z. Guo and Z. Wu, ChemSusChem, 2018, 11, 688–699 CrossRef PubMed.
  55. W. Xu, X. Li, C. Peng, G. Yang, Y. Cao, H. Wang, F. Peng and H. Yu, Appl. Catal., B, 2022, 303, 120910 CrossRef.
  56. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef PubMed.
  57. G. K. D. Joubert, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 1758 CrossRef.
  58. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed.
  59. H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Solid State, 1976, 13, 5188–5192 CrossRef.
  60. J. Li, F. Zeng, J. K. El-Demellawi, Q. Lin, S. Xi, J. Wu, J. Tang, X. Zhang, X. Liu and S. Tu, ACS Appl. Mater. Interfaces, 2022, 14, 45254–45262 CrossRef.
  61. H. Li, X. Lv, R. Li, X. Tao and Y. Zheng, J. Power Sources, 2022, 522, 231006 CrossRef.
  62. M. Ding, W. Chen, H. Xu, C. Lu, T. Lin, Z. Shen, H. Tao and K. Zhang, ACS Appl. Mater. Interfaces, 2020, 12, 9209–9218 CrossRef.
  63. X. Lv, X. Li, C. Yang, X. Ding, Y. Zhang, Y.-Z. Zheng, S. Li, X. Sun and X. Tao, Adv. Funct. Mater., 2020, 30, 1910830 CrossRef.
  64. J. Low, J. Yu, M. Jaroniec, S. Wageh and A. A. Al-Ghamdi, Adv. Mater., 2017, 29, 1601694 CrossRef PubMed.
  65. H. Zhang, H. Gu, X. Wang, L. Li, J. Zhang, S. Chang and W.-L. Dai, J. Colloid Interface Sci., 2022, 622, 539–548 CrossRef PubMed.

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