Abstract
Photocatalytic hydrogen production is a key pathway toward sustainable energy, driven by semiconductors that utilize sunlight for water splitting. This review highlights recent advances in material design, theoretical modeling, and data-driven discovery. Focus is given to visible-light-active semiconductors with optimal band gaps (1.8–2.4 eV), such as BiVO4, g-C3N4, and CdS, which enable efficient redox reactions. Hybrid architectures, including Pt-loaded TiO2 and CdS/ZnS core–shell systems, demonstrate hydrogen evolution rates exceeding 105 mol m−2 s−1. Upconversion nanomaterials based on rare-earth-doped fluorides extend light harvesting into the NIR, enhancing quantum yields when combined with quantum dots. Engineered heterojunctions and carbon-based 2D interfaces improve charge separation and suppress recombination. Thermodynamic parameters such as low overpotentials (<0.3 V) and high absorption coefficients (>105 cm−1) correlate with high catalytic efficiency. Time-dependent simulations and density functional theory (DFT) offer insights into structure–property relationships. Additionally, machine learning models expedite discovery by navigating complex compositional and structural spaces. While integrating theoretical, experimental, and AI-driven approaches, this review presents a framework for the rational design of scalable photocatalysts that meet future energy demands.
1 Introduction
Carbon saturation in the atmosphere has created an immerse climate change (Abbass et al. 2022; Corwin 2021; Karl et al. 2009; Sesana et al. 2021; Wheeler and Von Braun 2013) emanating from heavy industrial consumption of fossil fuel, which accounts to 76 % of greenhouse gas emissions (GHG). Although there are around 265 carbon capture project sites, only about 40 are currently operational (Statista 2024). Compared to the rapid progress in renewable energy technologies, carbon capture still lags behind in deployment and efficiency. Nevertheless, as scientists and engineers strive to eliminate the use of hydrocarbons for both pre and post combustion processes, hydrogen has emerged as a promising alternative source of clean energy.
The rising demand of hydrogen is intended to improve and develop the human index. For this reason, many techniques for the production of hydrogen (Abdin et al. 2020; Dawood et al. 2020; Yue et al. 2021) have been developed with concerns to reducing the energy intensity to hydrogen production without emitting GHG and at a low cost. Some of these hydrogen production resources, which are considered safe are from biomass, solar, and wind with several evolving semiconducting catalytic materials. The most contending economical hydrogen generation techniques utilizing visible light are driven from solar energy coupled with catalytic reactions to form photocatalytic processes. Though this processes is promising, many techniques still suffers a challenge as best performing technique. It is quite less complex to split water (Astruc 2020; Chen and Arnold 2020; Takata et al. 2020) into hydrogen (H2) at cathode and oxygen (O2) at anode from the coupling of solar (photons), and catalyst, known as the photocatalytic water splitting.
Apart from sourcing photons from the abundance of lights from solar, near-infrared region, and light emitting diodes. These photocatalytic materials (Chen et al. 2020; Shen et al. 2020; Wu et al. 2023) used for water splitting are predominately titanium dioxide (TiO2), ferric oxide (Fe2O3), tungsten trioxide (WO3), bismuth vanadate (BiVO4) and zinc oxide (ZnO) among others. However, these materials are generally considered semiconductors with several different band gaps, of which some could be heterogeneous, homogeneous, or plasmonic antenna-reactor structured type. TiO2 with a wide band gap absorbs photons from light spectrum of about 200 nm–800 nm, this material excited electrons transition from its valence band to conduction band thereby creating holes in valence. The electrons and holes generated by the photo effect move towards the catalyst surface and actively engage in the process of reducing and oxidizing the water molecules that are adsorbed (Albero et al. 2020; Karimi-Maleh et al. 2020; Ndia Ntone et al. 2024; Wang et al. 2021).
The fundamental engineering processes of photocatalytic quantum efficiency are mainly from light harvesting, surface adsorption capacity, charge separation, transport, and utilization; Figure 1 highlights the different methods to photon-induced green hydrogen production. That said, it is reported by the United States Department of Energy (DOE) as of 2020 that solar to hydrogen (STH) (Zhou et al. 2023) conversion amounts to 20 % and has an ultimate target of 25 % improvement (US Department of Energy 2024). Therefore, the continuous modeling of materials using the ab initio molecular dynamic principles and other predictive computational intelligence have been resorted to finding the best optimum structure for experimental photocatalytic processes to improving on light-to-hydrogen conversion (LTH).

Illustration of three light-driven water splitting systems: (a) photocatalytic (PC), (b) photoelectrochemical (PEC), and (c) photovoltaic–electrochemical (PV–EC), showing distinct mechanisms for solar hydrogen generation, including light absorption, charge separation, and redox reactions at active interfaces. Reproduced from (Bian et al. 2021), journal of energy chemistry, 57, 325–340 (2021), Elsevier, under the terms of the creative commons attribution 4.0 international license (CC BY 4.0).
The application of science and artificial intelligence (SciAI) (Xu et al. 2021; Zhang and Lu 2021) in material modeling for several hydrogen production methods, such as direct air electrolysis (DAE), anion exchange membrane (AEM), proton exchange membrane (PEM), biomass gasification, microbial electrolysis cells, GenHydro reactor system, photo-electrochemical catalysis (Ko et al. 2022; Ursua et al. 2011) (PEC) to mention but a few have gained tremendous traction in the clean energy industry. This intelligent predictive and assessment of preferred heterogeneous catalysis have by far reduced the recent complexities and time associated to synthesizing materials in the laboratories. Large volumes of dataset for water splitting photocatalytic processes either from density function theory (DFT) calculations or from experimental analysis can be trained using machine learning or deep learning algorithms. While this is feasible, efforts are still at the elementary stage in the industry.
The significance of the input parameter is as important as tweaking hyperparameters of the algorithms to optimize process parameters. Since data plays a critical role for design optimization, ScienceDirect shows a total of 145,382 photocatalytic results (ScienceDirect 2024). Sampling experimental or published data can include high degree of inconsistencies from different research groups with several process and measuring equipment, and this makes training and validation quite complex. For this reason, standardizing parameters for photocatalytic processes from any derived light source is a major discussion for the industry. That said, learning patterns of data with different architectures from regressions to convolutions (Isazawa and Cole 2023; Kumar and Singh 2021; Masood et al. 2019) are explored in this study under several light to hydrogen (LTH) descriptors, such as energy band gap, absorption co-efficient, and electron energy loss function.
Despite some data-driven inaccuracies for coupling artificial intelligence with material modeling efficiencies. This critical review takes a careful discussion on the evolution, recent development, challenges and perspectives of first principles molecular dynamics (Irawan et al. 2024), semiconductor materials and light sources, photocatalytic hydrogen production methods, machine and deep learning architectures. Figure 2 presents an informed framework and discussion to support photocatalytic experts of the recent intelligence to aid managerial decision making.

Schematic process of photocatalytic database from MD/DFT codes to machine and deep learning algorithms.
2 Ab initio molecular dynamics (AIMD) and density functional theory (DFT) approach
Spectroscopy provides a window into the experimental characterization of catalysts and their atomic-scale interactions with light. To complement these findings, computational methods are essential for probing electronic structures such as phonons, band gaps, and electron energy loss spectra. Modeling the molecular dynamics of photocatalytic reactions has spurred the development of algorithms capable of treating electrons as both particles and waves. A search on GitHub reveals over 1,100 molecular dynamics repositories (Leimkuhler and Matthews 2015), yet identifying robust, up-to-date implementations grounded in modern quantum mechanics remains a challenge. Notably, several actively maintained platforms support the evolving needs of quantum chemistry and physics, particularly those grounded in Molecular Dynamics (MD) and Density Functional Theory (DFT), as summarized in Table 1.
Summary of AIMD and DFT software tools, their applications, limitations, and distinctive computational features.
| Software/method | Applications | Limitations | Mechanistic highlights/discussion |
|---|---|---|---|
| AIMD | Models thermal and atomic motion; dynamic evolution of systems under irradiation or temperature | High computational cost; typically used for short time scales | Captures time-dependent atomic trajectories; complements DFT for exploring phonons, reaction pathways, and thermal stability |
| DFT (general) | Ground-state energy; electronic, magnetic, and optical properties; band structure and DOS | Underestimates band gaps; accuracy depends on XC functionals | Widely adopted due to balance of accuracy and efficiency; reliable for structural and electronic predictions |
| VASP | Periodic solids, interfaces, quantum MD, hybrid functional calculations | Commercial license; expensive; GPU demanding | Plane-wave-based; supports HSE, GW, and RPA corrections; used for surface and defect modeling |
| SIESTA/ASAP | Molecular + solid systems; electronic structure and MD; phonon/vibrational modes | Full version post-2003 not fully free; steep learning curve | Uses localized basis sets; integrates TDDFT, NEB, and optical transport; ASAP platform aids non-programmers |
| CASTEP | Band structures, phonons, optical spectra, surfaces, zeolites | GUI support limited unless paired with materials studio | Plane-wave DFT; specialized in difficult systems where experimental data is limited; good for solid-state physics |
| BigDFT | Large-scale systems; GPUs for performance boost; real-space DFT | Less community usage than VASP or QE | Uses daubechies wavelets; optimized for hybrid CPU-GPU architectures; good for open boundary systems |
| Quantum ESPRESSO | Planewave-based DFT and MD; wide plugin ecosystem | Requires command-line skills; GPU support partial | Widely used in academic research; modular, customizable, supports LDA, GGA, hybrid functionals |
| ABINIT | Full DFT suite; relaxation, MD, perturbation theory | Steep learning curve for advanced modules | Similar to QE; focus on pseudopotentials and response functions; supports GW and TDDFT |
| Q-chem/ORCA | Molecular systems; quantum chemistry, TDDFT, HF | Less suited for solids or extended periodic systems | Strong post-HF and hybrid DFT support; widely used in molecular spectroscopy |
| Mat3ra (GUI platform) | Cloud-based modeling interface; VASP, SIESTA, quantum ESPRESSO | Limited customization; cost for advanced features | Enables remote HPC execution; useful for students and non-programmers needing intuitive interface |
Ab initio MD simulates the thermodynamic evolution of atoms and molecules from first principles, while DFT focuses on determining the ground-state electronic properties and total energy for fixed atomic positions (GitHub 2024; Orio et al. 2009). DFT is now widely regarded as the leading method for modeling semiconductor materials in photocatalysis. Before physical synthesis or modification, structural, optical, magnetic, and electronic characteristics of photocatalysts can be reliably predicted via DFT, reducing experimental burdens. This is made possible by the foundational principles of quantum mechanics, which minimize reliance on empirical parameters.
Explorations across the periodic table have enabled researchers to compute minimum-energy configurations, known as ground states, where the principal quantum number n = 1. DFT codes often rely on self-consistent field (SCF) solutions from Kohn-Sham equations, incorporating cell relaxations, densities of states, wavefunctions, band structures, and phonons. Because of inaccuracies in Coulomb potentials, effective core potentials or pseudopotentials are introduced. These include Kohn-Sham exchange-correlation (XC) functionals (Cosmologic-service 2024), which capture the interactive electron energy. Commonly adopted functionals include Local Density Approximation (LDA) (Kohn and Sham 1965), Generalized Gradient Approximation (GGA) (Perdew 1986; Perdew et al. 1992, 1996), and hybrid forms like PBE0 (Hammer et al. 1999). These DFT methods, which employ ultrasoft or norm-conserving pseudopotentials, often demand substantial computational resources, requiring high-performance CPUs and GPUs. As shown in Figure 3, Genovese et al. (2009) demonstrated significant speedups using GPUs in hybrid CPU-GPU architectures.

The integration of GPUs with CPU cores for running DFT codes. The plot highlights GPU speedup trends for varying atom sets, emphasizing computational efficiency improvements when optimizing GPU–CPU configurations. Adapted with permission from (Genovese et al. 2009). The journal of chemical physics, 131(3), 034103 (2009). © 2009 AIP publishing. Licensed for reuse in reviews in chemical engineering, license number 6070241185826.
Their tests using BigDFT demonstrated consistent GPU acceleration across different atom counts and CPU-core distributions. For example, speedup remained stable for 68 atoms even with more cores, while gains plateaued at 32 cores for 128 atoms. These findings suggest that efficient configurations can be achieved with one GPU and a limited number of CPU cores. Accurately estimating the bandgap of semiconducting materials is essential for evaluating their photocatalytic viability, particularly for solar-driven hydrogen production. The bandgap determines both the wavelength of light that can be absorbed and whether the material possesses sufficient redox potential for water splitting. For overall photocatalytic activity, an optimal bandgap typically falls in the range of 1.6–2.4 eV (Walter et al. 2010), balancing light absorption and redox capability. However, standard DFT calculations often underestimate experimental bandgaps due to limitations in the exchange-correlation functionals.
For instance, the commonly used PBE functional underestimates the bandgap of anatase TiO2 (experimentally 3.2 eV) by more than 1 eV. This discrepancy stems from the self-interaction error and the lack of derivative discontinuity in GGA-type functionals (Perdew 1985). To mitigate this, hybrid functionals such as HSE03, HSE06, and PBE0 incorporate a fraction of exact Hartree–Fock exchange, improving the accuracy of bandgap predictions. HSE variants additionally apply range-separated screening, balancing computational efficiency with predictive power. While these approaches are more demanding computationally, they offer a more reliable estimate of the conduction and valence band edges, which are crucial for assessing HER/OER feasibility.
Sham and Schlüter (Sham and Schlüter 1983) introduced a formalism for evaluating the bandgap based on the Kohn–Sham eigenvalues, but this remains an approximation without quasiparticle corrections. More accurate methods such as the GW approximation or Δ SCF are often employed to refine band alignment, particularly for high-throughput screening. Thus, bandgap prediction requires careful choice of functional and methodology to avoid overestimating photocatalytic potential.
Scientific programming for materials modeling requires substantial skill, particularly when software must be compiled and executed in low-level languages. Currently, over 54 simulation tools exist, with ∼90 % developed in FORTRAN and others in C, Python, CUDA, or Perl. Many tools require command-line installation, which can deter users unfamiliar with programming. Proprietary software often provides user-friendly interfaces, whereas free software under LGPL, GPL, BSD, or MIT licenses typically lacks graphical environments. Among these, 26 are fully open-source, though licenses like MIT have minimal constraints. Within the scope of MD, DFT, and TDDFT simulations, free tools like BigDFT (Genovese et al. 2011), ABINIT (Gonze et al. 2009), Quantum ESPRESSO (Giannozzi et al. 2009), and legacy SIESTA (Soler et al. 2002) stand out. Note that post-2003 versions of SIESTA are no longer open-source.
Supporting open-access simulation software broadens research accessibility for scientists worldwide. Among the 54 tools, 16 operate under academic licenses. Notably, software like VASP (Hafner 2008), ASAP-SIESTA (ASAP-SIESTA 2024), CASTEP (Clark et al. 2005), Q-Chem (Shao et al. 2015), and ORCA (Neese 2012) support MD, DFT, and TDDFT.
Popular academic platforms include SIESTA, VASP, and CASTEP. SIESTA, compiled in FORTRAN 95, excels at large-scale electronic and molecular dynamics modeling, solving self-consistent Kohn-Sham equations. ASAP now provides a graphical platform for SIESTA, allowing simulations without programming. Capabilities include single-point energies, structural relaxations, molecular dynamics, phonons, optical responses, and transport.
CASTEP uses DFT with plane-wave pseudopotentials for quantum simulations in solid-state and chemical engineering applications. It is ideal for systems lacking empirical models. It supports surface, crystal, and defect studies.
VASP solves Kohn-Sham or Roothaan equations using plane-wave bases and PAW or Vanderbilt pseudopotentials. It enables quantum-level modeling of solids, polymers, and molecules. Post-DFT enhancements in VASP include HF exchange, hybrid functionals, and many-body corrections.
Most of these tools operate on local machines. ASAP-SIESTA supports remote execution, while Mat3ra (mat3ra 2024) enables browser-based, cloud-integrated simulations. A schematic of the discussed tools is shown in Figure 4.

Schematic flow of recent free and academic software licenses for molecular dynamics, density functional theory and time-dependent density functional theory calculations.
3 Evolution of light–driven photoreactions
Light-driven photoreactions are the cornerstone of photocatalytic water splitting, offering sustainable avenues to harness solar energy for hydrogen production. Over the years, the progression of light sources and materials has significantly shaped the field, particularly through advances in photon absorption efficiency and energy conversion pathways. This section delves into the evolution of light sources, ranging from solar and near-infrared (NIR) radiation to mercury discharge lamps and light-emitting diodes (LEDs), and their applications in photocatalysis.
3.1 Solar–NIR radiation
The utilization of solar energy in the photocatalytic splitting of water has been pivotal in advancing clean energy technologies. Solar energy is abundant, cost-free, and inherently sustainable, making it the most attractive light source for large-scale hydrogen production. Traditional photocatalytic materials, such as titanium dioxide (TiO2), were initially developed to absorb ultraviolet (UV) light, which comprises only about 4–5 % of the solar spectrum. However, as research advanced, it became evident that harnessing the visible and near-infrared (NIR) portions of the spectrum was critical for improving overall solar-to-hydrogen (STH) efficiency (Agrawal et al. 2024; Kumar et al. 2024; Patel et al. 2024; Priya et al. 2024). Modern solar simulators are of late considered a practical step to simulating electrochemical cells for water splitting in the laboratory; Figure 5 showcases a typical solar-simulator in the University College London for water splitting.

Solar simulator-driven photoelectrochemical water splitting, illustrating the experimental setup used to evaluate photocatalyst efficiency under simulated sunlight. Adapted with permission from the University College London website (University College London 2024).
Recent innovations in material science (Ansari and Sillanpaa 2021; Cao et al. 2022) have enabled the development of semiconductors and doped materials capable of extending light absorption into the NIR range (700–2,500 nm). Upconversion nanoparticles (UCNPs), for instance, have emerged as effective mediators that convert NIR photons into higher-energy visible photons. By integrating UCNPs with conventional photocatalysts such as TiO2 or WO3, researchers have achieved significant improvements in water-splitting efficiency under NIR irradiation.
Moreover, tandem systems employing multi-junction photocatalytic materials have demonstrated improved photon utilization by sequentially absorbing photons across the UV, visible, and NIR regions. For example, bismuth vanadate (BiVO4) coupled with a silicon-based photoelectrode has shown promise in maximizing STH conversion (Jin et al. 2022).
Beyond photon absorption, the photothermal effects (Qureshi et al. 2024) of NIR radiation have gained attention. Photothermal catalysis leverages localized heating effects induced by NIR photons to enhance reaction kinetics at the catalyst surface. Plasmonic nanomaterials, such as gold and silver nanoparticles, are often used to amplify these effects, leading to improved hydrogen evolution rates.
3.2 Mercury discharge lamps
Historically, mercury discharge lamps as displayed in Figure 6 played a pivotal role in the early stages of photocatalytic research. These lamps emit a broad spectrum of light, including UV, which was instrumental in validating the photocatalytic activity of materials like TiO2.

High-pressure mercury lamp marketed for its broad emission spectrum ranging from 200 nm to 800 nm, commonly used in photochemical experiments. Adapted with permission from the Shilpent website (Shilpent 2024).
Mercury discharge lamps offer high UV intensity (Broxtermann et al. 2017), making them suitable for studying UV-responsive photocatalysts. However, their inefficiency in energy consumption and environmental concerns related to mercury disposal have limited their modern-day applicability. Despite this, these lamps remain valuable tools as elucidated in Figure 6 for laboratory-scale experiments that require precise control over UV–vis light exposure (Horikoshi et al. 2019; Kubiak 2024; Meinhardova et al. 2023).
One notable application of mercury discharge lamps is their use in evaluating quantum efficiencies under controlled conditions. Researchers have used these lamps to identify reaction mechanisms by isolating the effects of UV-induced charge carriers. However, the lack of compatibility with visible or NIR-responsive photocatalysts has led to a gradual decline in their use, particularly as the field transitions to more sustainable and efficient light sources.
3.3 Light emitting diodes (LEDs)
The advent of light-emitting diodes (LEDs) (Bhattarai et al. 2024) has revolutionized photocatalysis by providing an energy-efficient, tunable, and environmentally friendly alternative to traditional light sources. LEDs offer narrow-band emission, enabling researchers to tailor light wavelengths to match the absorption spectra of specific photocatalysts.
LEDs consume significantly less power than mercury discharge lamps and have a longer operational lifespan. Additionally, their ability to emit light across the UV, visible, and NIR spectra makes them versatile for a wide range of photocatalytic applications. By utilizing LEDs with customizable wavelengths, researchers can optimize photocatalytic processes for various materials and reaction conditions. For example, blue and red LEDs are often paired with visible-light-responsive photocatalysts like BiVO4 and cadmium sulfide (CdS), respectively.
Recent studies have demonstrated the integration of LED arrays in scalable photocatalytic reactors as demonstrated in Figure 7. Such systems combine the precision of wavelength control with the ability to uniformly illuminate large reaction surfaces. Additionally, the pulsed operation of LEDs has been explored as a method to enhance charge separation in photocatalysts. Pulsed LED systems intermittently provide light to the reaction system, allowing time for photogenerated charge carriers to participate in redox reactions before recombining.

Schematic of the water splitting and storage process under white-LED irradiation, highlighting the photocatalytic flow enabled by modified g-C3N4 materials. Reproduced from (Tarighati Sareshkeh et al. 2023), scientific reports, 13, 15079 (2023), nature publishing group, under the terms of the creative commons attribution 4.0 international license (CC BY 4.0).
The evolution of light-driven photoreactions is closely tied to the development of photocatalytic materials optimized for specific light sources (Balarabe and Maity 2022; Khodadadian 2019). For instance, UV-responsive materials like TiO2 were initially the primary focus, but advancements in doping, heterojunction formation, and plasmonic effects have expanded the applicability of materials to the visible and NIR regions. Furthermore, hybrid light sources that combine solar simulators with LEDs have been employed to mimic real-world conditions, enabling more accurate assessments of photocatalytic performance.
3.4 Emerging trends in light sources
Emerging light sources such as perovskite-based emitters and laser-induced systems represent innovative directions in photocatalysis, as summarized in Table 2. Perovskite materials (Lim et al. 2019; Zhu et al. 2019), renowned for their tunable bandgaps and high quantum yields, are increasingly explored for hybrid configurations where they serve as both absorbers and emitters.
Summary of key light sources in photocatalysis, their applications, limitations, and distinctive characteristics.
| Light source/system | Applications | Limitations | Mechanistic highlights/discussion |
|---|---|---|---|
| Solar–NIR radiation | Large-scale water splitting; sustainable H2 generation; PEC optimization | Limited UV share (∼4–5 %); inefficiency without broadband absorbers | NIR-sensitive catalysts (UCNPs, plasmonics) extend utility; tandem and multijunction systems (BiVO4/Si) enhance STH conversion; photothermal catalysis improves kinetics |
| Mercury discharge lamps | UV photocatalyst validation (TiO2); quantum efficiency tests | High energy consumption; mercury toxicity; no visible/NIR activation | Emit broad spectrum (200–800 nm); enable mechanistic studies with stable UV intensity; declining use due to environmental concerns |
| Light-emitting diodes (LEDs) | Tunable narrow-band photocatalysis; LED-reactor arrays; pulsed illumination | Lower photon flux than solar or Hg lamps; limited scalability in early versions | Energy-efficient, long lifespan; customizable wavelengths allow precise photocatalyst matching; pulsed LEDs aid charge separation and minimize recombination losses |
| Hybrid light sources (solar simulators + LEDs) | Lab-scale PEC mimicking sunlight; realistic testing conditions | Equipment complexity; cost; calibration needs | Combine solar spectrum and spectral control; bridge gap between simulation and real-world illumination; often paired with advanced semiconductors |
| Emerging: perovskite light sources | Internal photon upconversion; light-absorbing–emitting hybrid role | Stability; toxicity (Pb); photobleaching risks | Tunable bandgaps, high QY; coupled with semiconductors for photonic hybrid systems; potential self-powered catalysts |
| Emerging: laser-based systems | Localized excitation; spatially controlled photocatalysis | Experimental stage; cost; thermal side effects | Enable site-selective activation; used in microreactor studies and quantum dot excitation; relevant for patterned PEC devices |
Their spectral flexibility enables precise matching with semiconductor photocatalysts. Meanwhile, wireless light sources such as laser-based systems are being employed to achieve spatially resolved excitation, allowing for selective activation of catalytic sites. Although still in developmental stages, these techniques offer promising opportunities for precision-controlled and patterned photocatalytic reactions.
4 Photocatalytic semiconductor materials
Semiconductors form the backbone of photocatalytic water splitting, with their ability to absorb photons and generate charge carriers (electrons and holes) crucial to the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER). Over the years, significant advancements have been made in the synthesis, doping, and modification of these materials to improve their efficiency, stability, and adaptability to various light sources. This section explores the characteristics, recent innovations, and challenges associated with essential semiconductor materials used in photocatalysis (Goodarzi et al. 2023).
Titanium dioxide is the most widely studied and utilized photocatalyst, owing to its stability, abundance, and non-toxicity. Its bandgap of ∼3.2 eV allows it to absorb ultraviolet (UV) light, which limits its utilization of the solar spectrum but makes it effective for controlled experiments. To overcome the limitations of TiO2’s UV specificity, researchers have focused on doping and surface modifications. Metal doping with elements such as Ag and Fe introduces intermediate energy states within the bandgap, enabling visible light absorption.
Non-metal doping using elements like nitrogen and sulfur shifts the bandgap edge by modifying the electronic structure, improving light absorption in the visible spectrum. Heterojunction formation, combining TiO2 with other semiconductors such as WO3 or ZnO, enhances charge separation and reduces recombination rates. Nanoscale TiO2 including nanotubes, nanowires, and nanoparticles, has been extensively studied due to its high surface area and enhanced photoreactivity. The synthesis of vertically aligned TiO2 nanotube arrays, for instance, has shown promise in improving electron transport efficiency (Miyoshi et al. 2018; Yang et al. 2024).
WO3 (Wang et al. 2012), with a bandgap of 2.6–2.8 eV, is a visible-light-responsive semiconductor known for its stability under acidic conditions. WO3’s ability to drive oxygen evolution reactions efficiently has made it an essential component in tandem systems. Coupling WO3 with hydrogen-evolution catalysts (Sarkar et al. 2024) such as Pt improves overall water-splitting efficiency. Plasmonic nanoparticles such as gold and silver integrated with WO3 create localized surface plasmon resonance (LSPR), boosting visible light absorption and promoting hot-electron transfer.
ZnO, similar to TiO2, is a UV-active semiconductor with a bandgap of ∼3.37 eV. While it shares many properties with TiO2, ZnO is more susceptible to photocorrosion, limiting its long-term applicability. Researchers have developed composite systems, combining ZnO with carbon-based materials such as graphene and carbon nanotubes to mitigate photocorrosion and enhance charge transport. ZnO is often paired with other semiconductors in binary systems to improve stability and broaden light absorption (Abou Zeid and Leprince-Wang 2024). For example, ZnO/CdS heterojunctions have demonstrated enhanced photocatalytic performance under visible light (Qi et al. 2023).
BiVO4 has gained prominence for its ability to absorb visible light (bandgap ∼2.4 eV) and its compatibility with water oxidation reactions (Tolod et al. 2020). Monoclinic BiVO4 exhibits superior photoactivity compared to tetragonal or scheelite structures. Modifications such as introducing oxygen vacancies or forming heterojunctions with materials like g-C3N4 have further improved its catalytic properties. BiVO4 is frequently used in tandem PEC cells, where it serves as a photoanode, complementing hydrogen-evolution photocathodes. This configuration maximizes solar-to-hydrogen conversion efficiency.
4.1 Emerging binary semiconductor materials
Binary semiconductor systems are gaining traction due to their synergistic properties. These materials combine the strengths of individual components, such as enhanced light absorption, improved charge separation, and stability (Meyer et al. 2012).
The combination of cadmium sulfide (CdS) and zinc sulfide (ZnS) in core–shell structures helps reduce photocorrosion and improve charge separation as demonstrated by Vamvasakis et al. (2023) in Figure 8. This system is particularly effective under visible light, achieving high hydrogen evolution rates (Jiang et al. 2016). As a metal-free semiconductor, graphitic carbon nitride (g-C3N4) offers a unique approach to photocatalysis, as in the case of (Wang et al. 2024a). Its bandgap of around 2.7 eV allows it to absorb visible light, and its compatibility with various co-catalysts makes it versatile. Coupling g-C3N4 with noble metals like platinum or ruthenium significantly enhances its photocatalytic activity (Vasilchenko et al. 2022). Perovskite materials, with their adjustable bandgaps and excellent light absorption properties, are emerging as promising candidates for high-efficiency photocatalysis. However, stability issues continue to be a significant challenge.

Demonstration of CdS/ZnS nanocomposites as a binary semiconductor system operating under visible light for enhanced photocatalytic hydrogen generation. Reproduced from (Vamvasakis et al. 2023), nanomaterials, 13(17), 2426 (2023), MDPI, under the terms of the creative commons attribution 4.0 international license (CC BY 4.0).
Developing semiconductor materials for photocatalysis still faces several challenges, despite progress. Achieving the right balance between light absorption and charge carrier movement in the bandgap remains a critical issue. Many semiconductors, especially those sensitive to visible light, are prone to photodamage or deterioration during reactions. The rapid recombination of the generated electrons and holes reduces overall efficiency, requiring innovative approaches to separate charges. Scaling up the production of high-performing semiconductors while maintaining consistency and cost-effectiveness is a significant obstacle (Villa et al. 2021).
Research is now concentrating on combining improved materials and methods to solve these problems. Using machine learning in material design allows for predictive modeling and optimization, which is speeding up the discovery of new semiconductor materials. The use of quantum dots as light-absorbing agents shows potential for expanding light absorption into the near-infrared region. Combining organic and inorganic materials in hybrid systems provides a way to achieve both high efficiency and stability in photocatalysts (Ma et al. 2023).
4.2 Hybrid quantum dots and upconversion materials
Hybrid quantum dots (QDs) and upconversion materials (Chen et al. 2022; Gui et al. 2024) represent cutting-edge advancements in the field of photocatalytic semiconductors. These materials are designed to overcome the limitations of conventional semiconductors, such as restricted light absorption and inefficient charge carrier dynamics. By combining quantum dots with upconversion nanoparticles, researchers have created hybrid systems capable of extending light absorption into the near-infrared (NIR) region and enhancing photocatalytic efficiency.
4.2.1 Quantum dots in photocatalysis
Quantum dots (Jacak et al. 2013) are tiny semiconductor particles that exhibit unique size-dependent optical and electronic properties. Their adjustable bandgaps allow for precise control over light absorption, making them highly effective for photocatalytic applications. These dots can absorb light across a wide range of the spectrum, including visible and near-infrared regions. Their small size reduces charge recombination rates, improving overall photocatalytic performance. Surface modifications with ligands or co-catalysts can enhance their interaction with water molecules and reaction intermediates. While cadmium-based quantum dots, such as CdS and CdSe, have been extensively studied for their high quantum efficiency (Gomes et al. 2011), concerns over toxicity have led to the development of alternative, lead-free (Leng et al. 2016) options like ZnS and CuInS2, which are more environmentally friendly.
4.2.2 Upconversion materials
Upconversion materials can convert low-energy near-infrared (NIR) light into higher-energy visible light, allowing the activation of photocatalysts that cannot be activated by NIR light alone. These materials typically consist of rare-earth-doped nanoparticles, such as NaYF4 doped with ytterbium (Yb3+) and erbium (Er3+) ions. The energy transfer upconversion (ETU) process involves the absorption of multiple low-energy photons, which leads to the emission of a single high-energy photon. The photon avalanche is a cascade of energy transfer events that amplifies the upconversion process. Upconversion nanoparticles are combined with semiconductors like TiO2, BiVO4, or WO3 to expand their activity into the NIR region. This hybrid system improves the utilization of the solar spectrum and enhances photocatalytic efficiency (Zhou et al. 2015).
4.2.3 Hybrid quantum dot–upconversion systems
Combining quantum dots and upconversion materials creates a synergistic system that leverages the strengths of both components. Upconversion nanoparticles act as a foundation, converting near-infrared light to visible light, which is then absorbed by the quantum dots. Alternatively, quantum dots and upconversion materials are integrated into a single nanostructure, enabling efficient energy transfer and reducing recombination losses. These hybrid systems effectively enhance light-driven water splitting reactions. Quantum dots capture high-energy photons for electron generation, while upconversion materials provide additional excitation under near-infrared light, improving the hydrogen evolution reaction. Furthermore, the improved charge separation and reduced recombination significantly boost the efficiency of the oxygen evolution reaction.
Hybrid QD-upconversion systems have great potential, but there are still challenges to overcome. The use of cadmium and lead in quantum dots raises environmental concerns, so non-toxic alternatives must be developed. Improving the efficiency of energy transfer between upconversion materials and quantum dots is crucial to minimize losses. Scaling up the production of these hybrid systems while maintaining consistency and performance is a significant obstacle (Doughan et al. 2015; Liu et al. 2018). The solution lies in developing non-toxic, widely available materials, exploring advanced fabrication methods like self-assembly to enhance scalability, and integrating machine learning to optimize the design and performance of hybrid systems.
4.3 Material rate kinetics, thermodynamic thresholds, and light absorption
The performance of photocatalytic systems hinges not only on material properties but also on their intrinsic kinetics and thermodynamic constraints. A critical benchmark in hydrogen evolution reactions (HER) is the overpotential (η), the excess potential required beyond the standard redox potential to drive water splitting. For effective H2 generation, the conduction band minimum (CBM) of the photocatalyst must lie more negative than the H+/H2 redox potential (0 V vs. NHE), while the valence band maximum (VBM) should be more positive than the O2/H2O redox potential (+1.23 V vs. NHE).
Wang et al. (2023b) highlighted that the rate of hydrogen evolution can be described by the Arrhenius-like relation
Furthermore, light absorption coefficients (α) govern the depth of photon penetration and hence the effective carrier generation. Li et al. (2025b) employed time-resolved spectroscopy to correlate high α values (>105 cm−1) in narrow bandgap semiconductors with superior internal quantum efficiency (IQE), especially in NIR-driven systems. The interplay of α, carrier diffusion length, and surface recombination velocity (S) ultimately controls the external quantum efficiency (EQE).
To guide material selection, Figure 9 illustrates a simplified schematic of the band positions, overpotentials, and light penetration profiles in representative TiO2, ZnO, BiVO4 and other systems shown in Table 3, highlighting their HER thermodynamics and optical constraints.

Schematic of band alignment, overpotential thresholds, and representative light absorption depths for selected photocatalysts.
Summary of popular photocatalytic materials, their applications, limitations, and distinguishing mechanisms.
| Material | Applications | Limitations | Mechanistic highlights/discussion |
|---|---|---|---|
| TiO2 | UV-driven HER/OER; PEC water splitting; dye degradation | Wide bandgap (∼3.2 eV); poor visible light utilization; recombination losses | Stable, nontoxic; enhanced via N/Fe doping, Pt loading, and heterojunctions with g-C3N4; surface engineering improves carrier separation and QE |
| ZnO | UV-photocatalysis; organic pollutant degradation; H2 production | Photocorrosion; instability; UV-only response | Similar to TiO2; surface defects (oxygen vacancies, zinc interstitials) critical; improved via doping, ZnO/CdS heterojunctions, and CNT/graphene composites |
| BiVO4 | Visible-light photoanode for PEC water oxidation | Short hole diffusion length; slow kinetics; surface recombination | Monoclinic phase is favored; hybrid interfaces with g-C3N4, rGO improve photocurrent; supports tandem systems |
| WO3 | Visible-light-driven OER; acidic stability | Low conduction band; slow charge transport without co-catalyst | Works with Pt, Au plasmonics; LSPR boosts visible absorption and hot-electron injection |
| CdS/ZnS | Visible-light HER; reduced photocorrosion | Cd toxicity; photocorrosion under strong light | Core–shell morphology enhances charge separation; improved via co-catalyst (Pt, MoS2) integration |
| g-C3N4 | Metal-free photocatalyst; visible-light H2 evolution | Low surface area; rapid e−/h+ recombination | Tunable structure; synergy with Ru, Pt, or TiO2; used in Z-schemes |
| Perovskites | Bandgap-tunable photocatalysts; high light absorption | Moisture sensitivity; long-term instability | Good optical tunability; stability enhanced via compositional tuning and encapsulation |
| Quantum dots (CdS, CuInS2) | Visible–NIR photocatalysis; tailored optical properties | Toxicity (Cd, Pb); limited scale-up protocols | Size-dependent bandgap tuning; surface ligands and doping improve activity; low recombination rates |
| Upconversion materials (NaYF4:Yb,Er) | Extend activation to NIR range; solar utilization | Weak emission intensity; upconversion inefficiency | ETU and avalanche mechanisms; work synergistically with TiO2, BiVO4 in hybrid systems |
| QD–upconversion hybrids | Broadband solar-to-H2 conversion; improved reaction kinetics | Integration complexity; scalability; energy loss at interface | Combines NIR absorption and visible reactivity; high e−/h+ separation; ML optimization under exploration |
5 Water splitting mechanisms
Photocatalytic water splitting represents a crucial process for sustainable hydrogen production, involving the decomposition of water molecules into hydrogen (H2) and oxygen (O2) under light irradiation. This section explores the fundamental mechanisms governing water splitting, focusing on photon absorption, charge separation and transport, and surface reactions. Special attention is given to the roles of photocatalysts, co-catalysts, and reaction interfaces in enhancing efficiency and addressing challenges such as charge recombination and catalyst degradation.
Photocatalysts absorb incident photons with energy equal to or greater than their bandgap, promoting electrons from the valence band (VB) to the conduction band (CB). This creates electron-hole pairs that drive the subsequent redox reactions. Materials with a suitable bandgap (1.8–3.0 eV) can efficiently utilize sunlight. Wide-bandgap semiconductors like TiO2 primarily absorb UV light, while doped and heterostructured materials extend absorption into the visible and NIR regions.
Efficient charge separation is essential for preventing recombination and ensuring the availability of electrons and holes for redox reactions. Excited electrons in the CB migrate to the photocatalyst surface to participate in the hydrogen evolution reaction (HER), while holes in the VB oxidize water molecules, driving the oxygen evolution reaction (OER). The rates of these reactions are governed not only by intrinsic material properties but also by the structure and chemistry of the exposed surface.
Recent advances have emphasized the role of surface terminations, crystallographic facets, and defects in modulating catalytic behavior. For example, anatase TiO2 with exposed (101) and (001) planes exhibits markedly different adsorption energies for HER/OER intermediates. To model such site-specific activity, slab-based DFT calculations are used to simulate surface-dependent reactivity. Key descriptors like the adsorption free energies of H*, OH* and OOH* provide insight into the thermodynamics of reaction intermediates.
To scale this surface-aware modeling, large-scale datasets from resources like the CatHub database (Winther et al. 2019) and the Open Catalyst Project Chanussot et al. 2021 provide standardized DFT results for surface reactions across diverse materials and facets. Machine learning models, particularly those employing message-passing neural networks (MPNNs), such as DimeNet, or SOAP-based kernel approaches, have been trained on these slab-based datasets to rapidly predict adsorption energies, identify active sites, and guide catalyst discovery (Stark et al. 2023; Tang et al. 2024).
These ML-enhanced frameworks accelerate photocatalyst screening across variations in surface composition, terminations, and dopants, thereby bridging bulk electronic properties with atomistic surface energetics. This hybrid DFT–ML strategy not only enables rapid evaluation of candidate materials but also supports experimental targeting of specific facet-exposed nanostructures with optimal photocatalytic performance.
Water molecules adsorbed on the photocatalyst surface undergo reduction and oxidation reactions (Odabasi Lee et al. 2023):
The overall photocatalytic efficiency depends on a combination of electronic band structure, surface energetics, charge carrier lifetime, and co-catalyst configuration. By incorporating both DFT-derived and ML-augmented surface descriptors, researchers can develop predictive pipelines for rational catalyst design, accelerating the path toward scalable hydrogen production technologies.
5.1 Role of co-catalysts
Co-catalysts play a pivotal role in enhancing the efficiency of photocatalytic water splitting by facilitating charge separation and catalyzing surface reactions (Tian et al. 2023). Noble metals like platinum (Pt) are highly effective hydrogen evolution co-Catalysts (HER) co-catalysts due to their low overpotential and high conductivity. However, cost and scarcity have driven the exploration of alternatives such as transition metal phosphides (Ni2P) and Sulfides (MoS2). Oxygen evolution co-catalysts (OER) are often the rate-limiting step in water splitting. Co-catalysts such as IrO2 and RuO2 are commonly used but are expensive. Emerging options include cobalt-based catalysts (Co3O4) and Perovskite Oxides such as LaNiO3 (Qi et al. 2023). illustrates in Figure 10a how a two-step photoexcitation under a co-catalyst occurs with an aqueous redox mediator enhances water splitting processes, whereas (H. Zhang et al. 2024) demonstrates energy diagram for PEC water splitting reaction in Figure 10b using an n-type semiconductor photoanode and a counter electrode immersed in an electrolyte.

Redox-enhancing designs for PEC-based water splitting using co-catalyst strategies (a) and carrier dynamics (b). (a) Illustration of a Z-scheme energy diagram with co-catalyst positioning to enhance solar water splitting efficiency. Reproduced from (Qi et al. 2022), nature communications, 13(1), 484 (2022), nature publishing group, under the terms of the creative commons attribution 4.0 international license (CC BY 4.0). (b) Photoelectrochemical (PEC) water splitting involves three main steps: (i) Light absorption to excite electrons, (ii) charge transport through the semiconductor, and (iii) surface redox reactions via charge injection. Adapted with permission from (H. Zhang et al. 2024), sustainable energy and fuels (2024), Royal Society of chemistry. (© 2024 Royal Society of chemistry. licensed for reuse in reviews in chemical engineering).
5.2 Heterojunctions in charge separation
Heterojunctions (Wang et al. 2024a) are interfaces between different semiconductors, designed to separate charges by creating an internal electric field. One common configuration is the Type-II Heterojunction, which aligns the conduction and valence bands of two semiconductors to enhance charge separation, such as TiO2 coupled with CdS. Another type is the Z-Scheme System, which mimics natural photosynthesis by using two semiconductors with complementary band structures, where electrons and holes from the two photocatalysts recombine, leaving highly energetic charge carriers for redox reactions, as seen in systems combining BiVO4 and g-C3N4, which exhibit superior charge separation and redox capabilities.
Despite significant advancements, several challenges remain in achieving efficient and scalable water splitting. These include charge recombination, which reduces efficiency, and can be addressed by using co-catalysts to capture charge carriers and designing nanostructures to shorten charge carrier pathways. Catalyst stability is another challenge, as photocatalysts often degrade due to photocorrosion or oxidation, but can be addressed with protective coatings and stable co-catalysts. Additionally, both the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) require high overpotentials, increasing energy consumption, so developing catalysts with lower overpotentials is crucial. Finally, optimizing surface properties to enhance water molecule adsorption and reaction rates is a vital area of focus.
5.3 Light absorption and charge dynamics
Photocatalytic water splitting begins with photon absorption and exciton generation. Semiconductors with band gaps in the ∼1.8–3.0 eV range (BiVO4, g-C3N4, CdS) efficiently harvest visible light and have band edge positions straddling the H+/H2 and O2/H2O redox potentials shown in Figure 11a Upon band-gap excitation, electrons in the conduction band (CB) and holes in the valence band (VB) must be rapidly separated and transported to reactive sites, or else recombination will squander the absorbed energy. Material nanostructuring and heterojunction design are key to promoting charge separation: for instance, Type-II semiconductor junctions (TiO2/CdS) and Z-scheme architectures (BiVO4/g-C3N4) create internal fields or stepwise excitation pathways that funnel electrons and holes to different materials. Efficient charge dynamics are evidenced by longer charge-carrier lifetimes and suppressed recombination rates, often achieved by incorporating co-catalysts or surface defect states that trap carriers and mediate surface reactions. In summary, an optimal photocatalyst balances strong light absorption with effective charge separation, ensuring that photogenerated electrons and holes reach the surface with sufficient energy to drive the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER).

Energy band alignment and reaction pathway schematics for photocatalytic water splitting. (a) Schematic representation of energy band alignment in a photocatalyst system. The conduction band (CB) and valence band (VB) positions are aligned relative to the hydrogen and oxygen redox potentials. Co-catalysts and heterojunctions facilitate directional charge separation, improving photocatalytic efficiency by reducing recombination and enhancing electron–hole migration across interfaces. (b) Schematic free energy diagram illustrating the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) pathways on a catalyst surface. The energy barriers for intermediate steps (such as H*, OH*, OOH*) are depicted, highlighting thermodynamic requirements for efficient water splitting in photocatalytic and electrocatalytic systems.
5.4 Surface thermodynamics and slab modeling
Once charge carriers reach the surface, water splitting proceeds via adsorbate-driven reaction steps whose energetics can be elucidated by density functional theory (DFT) slab models. Key intermediates include adsorbed hydrogen (H*) for HER and adsorbed oxygen species (O*, OH*) for OER, as depicted in Figure 11b. The ideal catalyst surface binds intermediates neither too weakly nor too strongly (Sabatier principle), which for HER translates to a hydrogen adsorption free energy ΔGH* ≈ 0 eV as the optimum. DFT calculations on Ni3Mo alloys, for example, reveal how surface orientation and composition influence these thermodynamics.
In alkaline HER (where the initial Volmer step involves water dissociation), the Ni3Mo(101) facet exhibits a low water dissociation barrier (∼0.50 eV) and a near-optimal hydrogen binding energy (ΔGH* ≈ −0.18 eV, close to that of Pt), outperforming other Ni3Mo facets in predicted HER activity (Pham et al. 2021). These slab-model insights indicate that the Ni3Mo(101) surface provides an ideal dual functionality: Mo sites facilitate H2O activation while Ni sites optimally bind the resulting H*.
More generally, ab initio thermodynamic analyses map out free-energy profiles for each elementary step (H2O dissociation, H* adsorption/desorption, OH* formation for OER), allowing identification of the rate-limiting steps and energetics. Such calculations have shown that OER typically faces higher intrinsic barriers than HER due to multi-electron transfer requirements, rationalizing why OER co-catalysts (IrO2, Co–Pi) are often needed. By constructing surface Pourbaix diagrams and calculating activation energies on different facets or defect sites, researchers can predict surface reactivity trends and guide catalyst optimization prior to experimentation.
5.5 Surface-specific modeling of reaction energetics
Building on the insights from slab modeling, surface-specific approaches increasingly address the limitations of uniform facet assumptions.
While this study focuses on hybrid DFT and QML methods for photocatalyst discovery, we acknowledge that site-dependent reaction energetics, particularly those arising from surface terminations, defects, and vacancies, play a critical role in realistic water-splitting catalysis.
In high-throughput screening, this challenge is addressed via two complementary approaches: (i) surface-specific DFT datasets, such as the Open Catalyst Project (OC20/OC22), which explicitly model adsorption energies on diverse facet geometries and defect sites, and (ii) transfer learning techniques, where models pre-trained on bulk crystal or slab datasets are fine-tuned using adsorption energy data. Graph neural networks (GNNs), such as DimeNet++ and GemNet, have been adapted to capture local coordination environments, enabling the prediction of adsorption energies with site sensitivity.
Furthermore, recent developments in QML include active site attention mechanisms that allow models to focus on catalytically relevant atomic sites. Integrating such models into the screening pipeline ensures surface specificity without needing prohibitively expensive full DFT treatment for every site.
These models not only enhance the physical realism of catalyst screening but also serve as foundational tools for the DFT–ML workflows discussed next.
5.6 DFT–ML workflows for adsorption and energetics
Increasingly, high-throughput ab initio calculations are being coupled with machine learning (ML) to accelerate catalyst discovery as represented in Figure 12a. In this approach, DFT computations provide a database of structures, adsorption energies, and reaction barriers, which train ML models to predict energetics for unexplored materials. For instance, Chun et al. (2021) integrated a DFT-derived dataset with ML to screen ternary alloy electrocatalysts beyond the reach of brute-force DFT alone.

Illustration of machine learning strategies (top) and theoretical-experimental alloy design workflows (bottom) in catalysis. (a) Schematic diagram of ternary alloy configuration search, theoretical predictions, and experimental validation for the oxygen reduction reaction (ORR). Reproduced from (Chun et al. 2021), chem catalysis 1(4), 855–869 (2021), Elsevier, under the terms of the creative commons attribution 4.0 international license (CC BY 4.0). (b) An illustrative overview of computational paradigms for machine learning, including data-driven modeling and feature extraction strategies in materials science. Reproduced from (Li et al. 2022b), npj computational materials, 8, 127 (2022), under the terms of the creative commons attribution 4.0 international license (CC BY 4.0).
Their data-driven search identified an optimal PtFeCu composition, which was experimentally validated to exhibit ∼3-fold higher ORR mass activity than Pt/C (0.67 A mg−1 vs. 0.22 A mg−1) with superior durability. Although focused on the oxygen reduction reaction, this study demonstrates the power of ML-guided exploration of large compositional spaces relevant to water-splitting co-catalysts and electrodes. In the context of HER, Abraham et al. (2023) developed a multi-step ML workflow to predict hydrogen adsorption free energies on 2D MXene catalysts. They calculated ΔGH* for 1,125 candidate MXenes via DFT and trained a gradient boosting regressor to rapidly predict ΔGH* for thousands of others, achieving a low mean absolute error of ∼ 0.36 eV. This data-driven model pinpointed trends in composition: H binding on MXenes was most favorable when H adsorbed atop an outer metal site on O-terminated MXenes containing Nb, Mo, or Cr, with several predicted candidates showing ΔGH* near 0 eV (even surpassing Pt’s activity).
Likewise, single-atom catalyst (SAC) screening has benefited from DFT–ML synergies. Jyothirmai et al. (2023) e explored a series of metal and non-metal dopants on a g-C3N4 support, using DFT to evaluate H adsorption and then training ML models to generalize across the periodic table. Notably, they identified B, Mn, and Co single-atom dopants on g-C3N4 as top performers for HER, and among various algorithms tested, support vector regression yielded the best predictive accuracy in that study.
These examples highlight how DFT–ML workflows can rapidly screen vast design spaces (material compositions, surface facets, and ligands), deliver quantitative energetics (adsorption free energies within a few tenths of an eV of DFT values), and discover novel catalyst candidates with desirable thermodynamics. Such workflows typically involve feature engineering (to capture chemical descriptors like electronegativities, d-band centers), model training/validation (often requiring hyperparameter optimization and cross-validation for reliability), and uncertainty quantification to flag predictions for confirmatory DFT or experimental tests.
5.7 Integrated mechanistic perspectives
A comprehensive mechanism of photocatalytic water splitting emerges by integrating the above aspects, light absorption, charge carrier dynamics, surface reaction thermodynamics, and data-driven catalyst optimization, into a unified framework. In practical systems, these factors are interdependent: the semiconductor’s band alignment dictates the driving force for surface reactions, while the catalyst’s surface kinetics (a low ΔGH* or OER overpotential) dictates the required charge carrier energy and density.
Advanced studies increasingly emphasize this interplay. For example, co-catalyst nanoparticles (Pt, NiOx) are chosen not only for favorable HER/OER energetics but also for their ability to siphon electrons or holes efficiently from the light absorber, bridging the gap between charge dynamics and surface catalysis. Likewise, facet engineering in photocatalysts (exposing high-activity crystal faces) must go hand-in-hand with enhancing photon absorption in those regions (via light trapping or plasmonic enhancement) to truly leverage the facet’s catalytic potential.
From a modeling perspective, multi-scale simulations are being developed that couple photogenerated carrier transport models with surface reaction kinetics, providing an integrated picture of how a given material’s optoelectronic properties and surface chemistry collectively determine overall hydrogen output. This holistic view is crucial for guiding next-generation designs: it suggests, for instance, that simply having an ideal band gap is insufficient if surface reaction barriers remain high, and vice versa. Researchers are therefore pursuing synergistic improvements, tuning electronic structure (through doping and, phase junctions) and catalytic interface (through atomically dispersed active sites and, strain engineering) concurrently.
Future high-impact directions include dual-function materials that simultaneously optimize light harvesting and catalysis (plasmonic photocatalysts that generate “hot” carriers to lower reaction barriers) and closed-loop discovery platforms where experimental feedback refines DFT–ML models to account for real-world factors like kinetics and stability. In summary, an integrated mechanistic understanding, spanning photon to molecule, is being established as the foundation for breakthroughs in photocatalytic hydrogen production.
5.8 Advanced mechanisms and emerging trends
Plasmonic nanoparticles, such as gold (Au) and silver (Ag) (Lee and El-Sayed 2006; Link et al. 1999), enhance light absorption through localized surface plasmon resonance (LSPR) (Sarina et al. 2013). This generates hot electrons that participate in redox reactions. Plasmonic photocatalysis is particularly promising for NIR-responsive systems (Lee et al. 2024). Catalysts capable of performing both HER and OER on the same surface are being developed to simplify reaction systems and reduce costs. For instance, MoS2 modified with Ni(OH)2 has shown dual functionality (Subramanian et al. 2018).
Photothermal catalysis utilizes localized heating effects to accelerate reaction kinetics. This approach is often combined with plasmonic systems to enhance efficiency. To address current issues and make the technology commercially viable, research should concentrate on incorporating AI and machine learning to use predictive algorithms for finding ideal catalyst designs and reaction conditions, exploring abundant materials and scalable production methods to develop cost-effective catalysts, and designing multi-photon systems that can enhance light absorption.
These emerging strategies are reflected in recent experimental advances that demonstrate the practical outcomes of optimized photocatalytic systems. Table 4 provides a comparative summary of recent experimental efforts in binary photocatalytic systems for hydrogen evolution, highlighting variations in H2 yield, quantum efficiency (QE), and light source. These results underscore the role of material composition, heterojunction design, and co-catalyst integration in enhancing water-splitting performance.
Experimental performance summary of selected binary photocatalysts for H2 production.
| Reference | Photocatalyst | H2 yield | QE/IPCE | Light source |
|---|---|---|---|---|
| Wang et al. (2023a) | PtCu/TiO2 (sandwich) | 476.8 mmol g−1 h−1 | 99.2 % @ 410 nm | Visible LED |
| Takata et al. (2020) | Al:SrTiO3 + Rh/Cr2O3 + CoOOH | 253 μmol h−1 | 96 % @ 350–360 nm | UV |
| Li et al. (2022a) | COF nanosheets + Pt | 29.1 mmol g−1 h−1 | 82.6 % @ 450 nm | Visible |
| Zhang et al. (2023) | CdTe/V–In2S3 QDs | 12.1 mmol g−1 h−1 | 114 % @ 350 nm | UV |
| Q. Zhang et al. (2024) | Pt/g-C3N4 nanosheets (single atoms) | 336.8 μmol h−1 | 13.5 % @ 405 nm | Visible (λ > 400 nm) |
| Wang et al. (2019) | ZnO/CdS heterostructure (1:1, no Pt) | 1,040 μmol g−1 h−1 | N/A | Visible (λ > 420 nm) |
| Pihosh et al. (2015) | WO3-NRs/BiVO4 + CoPi (core-shell) | 6.72 mA cm−2 | ≈90 % (plateau ∼516 nm) | Simulated sunlight (AM 1.5G) |
For instance, PtCu/TiO2 and COF–Pt architectures show superior yield under visible light, while Al:SrTiO3-based systems excel under UV. This tabulated evidence reinforces the mechanistic discussions in Sections 5.1–5.3 by contextualizing how band structure engineering, charge separation strategies, and light harvesting directly translate to experimental hydrogen production efficiency.
6 Recent advances in machine learning for photocatalyst design
Machine learning (ML) is emerging as a transformative tool in photocatalytic water splitting research, enabling accelerated discovery and optimization of materials by learning from large datasets of experiments and simulations (Campos et al. 2024; Li et al. 2021; Wayo et al. 2023, 2024). Modern ML approaches can predict key properties (band gaps, stability, reaction rates) and identify patterns that would be difficult to discern via traditional trial-and-error. This section highlights recent advancements in applying both classical and quantum ML algorithms to photocatalyst design, with an emphasis on data handling, model interpretability, and integration with first-principles methods like Density Functional Theory (DFT) and ab initio molecular dynamics (AIMD).
Effective use of ML in materials science requires careful data preprocessing and noise management. Photocatalysis datasets often come from diverse sources with inconsistent experimental conditions, leading to significant variability. For instance, differences in light source intensity, solvent, or measurement techniques between labs can introduce noise that hinders model generalization. To address this, researchers are standardizing data collection and curation: in one study, Li et al. (2022b) measured hydrogen evolution for 572 organic molecules under identical conditions using a high-throughput setup, yielding a consistent training set for ML modeling (Figure 12b). Data preprocessing steps such as feature normalization (scaling inputs to comparable ranges) and outlier detection/removal are now standard to ensure that no single feature or erroneous data point dominates the training process.
In addition, data augmentation techniques are employed to expand limited datasets, for example, generating synthetic data points via DFT calculations or perturbing known structures, which can improve model robustness (Xu et al. 2023). Cross-validation (k-fold or leave-one-out) is used to quantify model performance under data noise and prevent overfitting, ensuring that the learned correlations reflect true physical trends rather than experimental scatter. These practices, coupled with open data initiatives (such as the Materials Project and Open Catalyst Project), are gradually mitigating the issue of data scarcity and inconsistency in photocatalysis ML research.
6.1 Classical machine learning in photocatalyst design
Classical (non-quantum) ML algorithms have become integral to accelerating photocatalyst discovery. Regression models predict continuous material properties critical to photocatalysis (Abraham et al. 2024; Arabaci et al. 2025; Wang et al. 2025). For example, linear and polynomial regression can provide baseline fits for how a dopant concentration or particle size affects the hydrogen evolution rate. More powerfully, non-linear regressors like Support Vector Regression (SVR) and Gaussian Process Regression (GPR) capture complex relationships such as how co-catalyst addition influences the band gap or surface charge transfer efficiency (Kumar and Singh 2021).
GPR, in particular, is valued for providing uncertainty estimates along with predictions, flagging which predicted photocatalyst performance values are less certain and may need experimental validation. Recent advances in classical machine learning (ML) have significantly enhanced photocatalytic materials discovery and optimization. Regression models predict key continuous properties such as bandgaps, reaction rates, and absorption coefficients. Linear regression provides baseline insights, while Support Vector Regression (SVR) is effective for capturing non-linear dopant-bandgap relationships (Awad and Khanna 2015). Gaussian Process Regression (GPR) enables uncertainty-aware predictions for unexplored material spaces (Deringer et al. 2021). Classification methods such as Random Forest (RF) (Breiman 2001; Rigatti 2017) and K-Nearest Neighbors (KNN) (Al-Dahidi et al. 2024; Javed et al. 2024; Peterson 2009) support activity screening, while Decision Trees aid interpretability (Kowsari et al. 2019).
Classification algorithms help screen materials by categorizing them (“active” vs “inactive” photocatalysts). Random Forest (RF) classifiers have been used to mine feature importance from large materials datasets, for instance, identifying which elemental or structural descriptors most strongly distinguish highly active photocatalysts. Decision trees and ensemble methods (Gradient Boosted Trees, Extreme Random Trees) produce human-interpretable decision rules, useful for quick heuristic screening. Clustering methods like k-Nearest Neighbors (KNN) can group similar materials and have been applied to discover clusters of semiconductors with comparable optoelectronic properties (Barik and Woods 2024; Lu et al. 2024). In a recent work, a suite of classifiers (KNN, RF, SVM, neural networks) was trained on molecular descriptors of photocatalysts and achieved >87 % accuracy in distinguishing high versus low H2-generation performers. Such models act as a filtration step to rapidly eliminate unpromising candidates from experimental consideration.
Deep learning methods are increasingly adopted as dataset sizes grow. Feed-forward neural networks (FNNs) and deeper multilayer perceptrons have been trained to predict photocatalytic efficiencies from rich feature sets (DFT-computed attributes, synthesis conditions), often outperforming simpler models in capture of non-linear effects. Convolutional Neural Networks (CNNs), originally developed for image recognition, have been repurposed to analyze material imaging data (SEM or TEM images of catalysts) and even diffraction or spectral patterns, automatically extracting microstructural features correlated with activity. Recurrent Neural Networks (RNNs) have seen use in analyzing time-series data such as real-time reaction kinetics, helping to model photocatalytic reaction progress under varying illumination or reactant feed conditions. Furthermore, autoencoders (unsupervised deep models) have been applied to reduce the dimensionality of materials data, for example, encoding a large set of candidate material compositions or structures into a smaller set of latent variables, which can then be explored to find novel candidates with similar latent features to known good photocatalysts.
Deep learning models expand ML capabilities. Feedforward Neural Networks (FNNs) predict electronic and photocatalytic properties (Das et al. 2024; Glorot and Bengio 2010; Yu and Cao 2025 ), and Recurrent Neural Networks (RNNs) analyze time-dependent reaction kinetics (Medsker and Jain 2001; Mienye et al. 2024). Autoencoders help extract latent structure-property relations, while Generative Adversarial Networks (GANs) and Variational Autoencoders (VAEs) (Kao et al. 2023; Mihandoost et al. 2024; Pakornchote et al. 2024; Yoon et al. 2024) drive inverse design. GANs have predicted novel materials with optimized bandgaps from high-throughput DFT datasets (Hu et al. 2022), and VAEs have suggested stable configurations for defect-tolerant structures.
Generative models are pushing the frontier of materials discovery by creating new hypothetical materials with targeted properties. Generative Adversarial Networks (GANs) have been trained on databases of semiconductor structures and their band gaps so that the generator network can propose new crystal structures expected to have optimal band gaps or stability for water splitting (Li et al. 2021). In one case, a GAN learned from thousands of DFT-evaluated compounds and was able to output candidate semiconductors with predicted band gaps in the 2.0–2.5 eV range (ideal for visible-light absorption) that were not present in the training set. Variational Autoencoders (VAEs) similarly have been used to explore chemical compositional space: by sampling the latent space, researchers generated novel alloy compositions predicted to be stable and photo-active, guiding experimental synthesis efforts. A recent study by Hu et al. (2022) combined high-throughput DFT with a GAN to successfully propose oxide photocatalysts with improved visible-light absorption (Kumar and Singh 2021).These generative models, while still nascent in application, hold promise for inversely designing photocatalysts by specifying desired performance metrics and letting the model “imagine” candidates to meet them.
Reinforcement learning (RL) techniques are being explored to optimize photocatalytic processes and synthesis in a closed-loop manner. In RL, an agent algorithm iteratively adjusts variables (actions) in response to feedback (rewards). For instance, researchers have implemented Q-learning and policy-gradient algorithms to dynamically tune reaction conditions (light intensity, pH, catalyst loading) to maximize hydrogen output. In one demonstration, an RL agent controlling a simulated photocatalytic reactor learned to maximize the hydrogen evolution rate by favoring a lower pH and higher light flux, balancing the trade-offs between reaction kinetics and catalyst stability.
Robotic experimentation platforms have also started using RL to guide the synthesis of photocatalysts: the agent proposes synthesis parameters (calcination temperature, dopant concentration), the experiment is performed, and the measured activity is fed back to the agent to inform the next suggestion. This closed-loop approach was shown to discover a better-performing co-catalyst deposition method for TiO2 in fewer iterations than a human-guided trial-and-error search (Haghshenas et al. 2024; Li et al. 2025a). While still emerging, RL offers a path toward self-driving labs for photocatalyst optimization. Reinforcement Learning (RL) is emerging for dynamic photocatalytic process optimization. RL agents adapt light intensity, catalyst concentration, and temperature for optimal H2 output. Q-Learning has been applied to hydrogen yield maximization (Watkins and Dayan 1992; Zhou et al. 2024 ), while Deep Deterministic Policy Gradient (DDPG) can tune continuous variables in multi-parameter systems (Xie et al. 2024).
Case studies and model interpretability: A growing emphasis in ML-assisted materials research is on interpreting the models to extract human-understandable insights. Black-box models can sometimes identify correlations without offering reasoning, so tools like SHapley Additive exPlanations (SHAP) and feature importance analysis are applied to decipher what the models have learned. For example, Kumar and Singh (2021) developed an interpretable ML model to screen 3,099 candidate 2D photocatalysts (metals with different ligand environments). By using SHAP values, they found that descriptors capturing differences in chemical hardness and electronegativity between the metal and ligands were among the most influential for predicting stability.
This feature attribution aligned with chemical intuition: highly stable compounds followed the Hard–Soft Acid–Base principle, where hard metal cations bond strongly with hard anions and vice versa. The SHAP analysis in their study highlighted, for instance, that a larger difference in electronegativity (
These predictions provide concrete targets for experimental synthesis and testing under real solar illumination. Indeed, bridging computation to experiment is a key validation step: an AI-driven screening is only truly successful when a predicted material is fabricated and demonstrates the anticipated performance in actual water-splitting tests. While many ML predictions await experimental verification, a few successes are reported; for example, Fe-doped graphitic carbon nitride (Fe/g-C3N4) was recently optimized via an ML model and then synthesized, achieving improved H2 evolution under sunlight consistent with the model’s forecast. Showcasing such end-to-end validations (prediction → fabrication → testing) will be crucial to build confidence in AI-designed photocatalysts.
Despite their promise, classical ML models face several challenges in photocatalysis applications. One major issue is the “small data” problem, high-quality experimental data are often limited due to the cost and time of measurements. Traditional ML algorithms typically excel with big data; with small datasets they risk overfitting. Researchers are tackling this by incorporating domain knowledge and advanced strategies: transfer learning (pre-training models on large related datasets like general inorganic compounds, then fine-tuning on specific photocatalyst data) and active learning (iteratively picking the most informative new experiments to perform, based on model uncertainty) have shown success in effectively using limited data. Another challenge is the integration of physical constraints.
Purely data-driven models might violate physical laws (predicting an impossible band structure); to prevent this, new physics-informed ML approaches embed constraints or known equations (like mass balance, thermodynamic limits) into the learning process. Model interpretability is also crucial: catalyst scientists are understandably hesitant to trust a black-box prediction without understanding the rationale. Interpretable models like decision trees or simpler linear models with physically meaningful descriptors are sometimes favored, or post-hoc explanation tools (like SHAP and LIME) are applied to complex models as illustrated above.
Finally, extrapolative predictions (predicting far outside the training domain) remain risky; an ML model trained on, say, oxide semiconductors may not reliably generalize to a novel metal-organic photocatalyst. This necessitates caution and often a human expert in the loop to assess whether a model’s prediction for an unprecedented material is sensible. In summary, classical ML has become a powerful complement to theoretical and experimental techniques in photocatalysis, especially when coupled with rigorous data handling and interpretability to ensure that the insights derived are physically meaningful and practically actionable.
The integration of experimental insights with machine learning (ML) architectures offers a powerful framework for materials discovery. Table 4, originally presented in the water splitting section, complements this discussion by providing empirical benchmarks, such as H2 yield and quantum efficiency, that serve as target variables for ML training. These metrics can be used in supervised regression and classification models to guide photocatalyst selection and optimization.
Additionally, Table 5, provides a detailed comparison between theoretically predicted and experimentally measured values for critical descriptors, band gaps, overpotentials, and H2 evolution rates, highlighting the discrepancies and uncertainties inherent in simulations. These gaps underscore the importance of coupling DFT predictions with experimental validation, which can be further bridged by ML models trained on such hybrid datasets. When used alongside the DFT–AI strategies outlined in Table 6, this tripartite integration (experiment–theory–AI) forms a closed-loop system for accelerating photocatalyst discovery. It enables not only predictive modeling but also uncertainty quantification and explainable recommendations, advancing interpretable, data-driven materials science.
Comparison of theoretical versus experimental band gaps, overpotentials, and hydrogen evolution rates for selected photocatalyst materials. H2 evolution rates are reported in mol m−2 s−1, normalized to the catalyst illumination area.
| Photocatalyst | Pred. Eg (eV) | Exp. Eg (eV) | Pred. overpot. (V) | Exp. overpot. (V) | H2 rate (pred/exp) |
|---|---|---|---|---|---|
| TiO2 (anatase) | 2.10 ( Zhang et al. 2016) | 3.20 (Zhang et al. 2016) | ∼0.5 (Stojkovic et al. 2024) | ∼1.67 (Sakar et al. 2019) | ∼3.6 × 10−4/∼2.8 × 10−5 (Haghshenas et al. 2024) |
| ZnO | 0.7–0.9 (Erhart et al. 2006) | 3.37 (Almeida et al. 2019) | ∼+0.5 (insuff.) (Erhart et al. 2006) | >1.8 (Nisar et al. 2025) | ∼2 × 10−5/<10−8 |
| BiVO4 (monoclinic) | 3.23 (Di Liberto et al. 2019) | 2.40 | 0 (HER limit) | ∼0.2 (bias needed) | ∼3.9 × 10−4/∼5 × 10−8 |
| CdS/ZnS | 2.06–2.25 (Zhao et al. 2025) | ∼2.29 | ∼0.31–0.44 | Lower HER | ∼4.6 × higher than CdS/14.44 mmol h−1 g−1 (Huang et al. 2023) |
| BiVO4/WO3 | ∼2.8 (Pihosh et al. 2015) | ∼2.4/2.5 | 0 V external | ∼0.21 V | ∼107 μmol h−1 cm−2/∼102 μmol h−1 cm−2 |
| QD–upconversiona | N/A | N/A | N/A | N/A | Up to 5 × 10−4/∼7.8 × 10−5 |
-
aQuantum-dot sensitizer with upconversion nanocrystal hybrid. This multi-photon system has no single band gap; sub-band-gap IR photons (980 nm, ∼1.27 eV) are upconverted to higher-energy photons absorbed by the quantum dot (Gao et al. 2023). Such hybrids broaden the absorption spectrum (UV–vis–NIR) and can enhance H2 yield, though performance depends on energy-transfer efficiency and specific band alignment.
Advances in DFT-AI integration for photocatalytic design for water splitting.
| Reference | DFT-AI integration | Materials investigated | Light source | ML architecture | Dataset | Performance metrics | Software (DFT/ML) |
|---|---|---|---|---|---|---|---|
| Huo et al. (2024) | Feature-assisted ML for band gap prediction | Binary semiconductors (1,208) | Not specified | SVR, RF, GBDT, SISSO | High-throughput DFT (PBE functional) | RMSE = 0.361 eV, R2 = 0.965 | VASP/Scikit-learn |
| Wang et al. (2024b) | ML-assisted screening for 2D materials | 2D materials (316,505) | Solar spectrum | ANN, RF, XGB | V2DB | RMSE <0.4 eV (band gaps) | Quantum ESPRESSO/Python |
| Jyothirmai et al. (2024) | Catalyst screening | g-C3N4/TMD heterostructures | Not applicable | Random forest regression (RFR) | Calculated gibbs free energy | MAE | VASP/Scikit-learn |
| Elbaz and Caspary Toroker (2024) | Band gap prediction | Spinel oxides (AB2O4) | Not specified | KRR, SVR, RF | Simulated spinel dataset | RMSE: 0.02 eV; conductivity MAE: 5 % | NEGF, GPAW/Scikit-learn |
| Sabagh Moeini et al. (2024) | Hybrid ML for band gap prediction | Low-symmetry perovskites | Not specified | SVR, RFR, GBR, XGBoost | CMR dataset (1,984 samples) | MAE <0.1 eV | HSE06/XGBoost |
| Oh et al. (2024) | Small dataset ML for band engineering | ZnTe-based alloys | Solar spectrum | SISSO + α-method | Custom ZnTe alloy database | RMSE: 0.1 eV | Quantum ESPRESSO/SISSO |
| Liu et al. (2022) | LightGBM for degradation rate prediction | Doped TiO2 (Ag, N, Cd) | UV-visible (254–600 nm) | LightGBM | Experimental dataset (760 points) | R2 = 92.8 % | Jupyter/Scikit-learn |
| Gladkikh et al. (2020) | KRR and extremely randomized trees | ABX3 perovskites | Not specified | KRR | HSE06 bandgap data (199) | RMSE = 0.3 eV | GPAW/Python |
| Ren et al. (2020) | GPR and sobol’ sensitivity analysis | TiO2 with Pt co-catalysts | Solar simulation | GPR | CFD + experimental data | Yield optimization | CFD tools/Python |
| Masood et al. (2023) | ML-accelerated DFT for material discovery | TiO2, CdS, WO3, g-C3N4 | Visible light | Random forest | Experimental + DFT outputs | R2 = 0.95 | VASP/Scikit-learn |
| Wexler et al. (2018) | ML for HER descriptor discovery | Ni2P with nonmetal dopants | Not specified | Regularized random forest | DFT structural data | HER ΔG_H = −0.11 eV | Quantum ESPRESSO/R (caret) |
| Li et al. (2018) | Predictive ML for stability | Ternary oxides | Not applicable | KRR, ET | ICSD | RMSE, MAE | VASP/PyCaret |
| Pereira et al. (2017) | ML for HOMO and LUMO prediction | Organic molecules (111,725) | Not applicable | Random forest | DFT HOMO/LUMO energies | MAE = 0.15 eV (HOMO), 0.16 eV (LUMO) | GAMESS/R |
6.2 Quantum machine learning (QML) in photocatalysis
Quantum machine learning combines principles of quantum computing with ML algorithms, offering potential speed-ups and new capabilities for materials design (Biamonte et al. 2017; Cong et al. 2019; Huang et al. 2021; Jayan and Babu 2024; Kao et al. 2023; Steane 1998; Vedavyasa and Kumar 2024). In the context of photocatalytic water splitting, QML is an exciting emerging area aimed at improving the accuracy of simulations and the efficiency of exploring candidate materials. The appeal of QML lies in the unique features of quantum computing: qubits can exist in superposition and become entangled, enabling certain computations to scale more efficiently than on classical bits. For materials problems, this means a quantum machine learning model might handle the quantum-mechanical complexity of catalyst behavior (such as electron correlation or tunneling effects) more naturally than a classical model.
One direction of QML research is using quantum computers to accelerate quantum chemistry calculations. Quantum algorithms (like the Variational Quantum Eigensolver) can potentially find ground-state energies or reaction barriers faster than classical quantum chemistry, which could allow rapid evaluation of photocatalyst properties within an ML loop Cerezo et al. (2021). For example, a quantum neural network or quantum kernel model could be trained to predict a material’s band gap or exciton binding energy by implicitly processing quantum states of that material.
Quantum Support Vector Machines (QSVMs) use quantum-enhanced feature spaces to perform classification of materials: initial demonstrations have shown that QSVMs can categorize crystal structures or material phases with accuracy comparable to classical SVMs while using fewer training samples. In one benchmark study, a QSVM was applied to classify transition-metal compounds and achieved ∼99 % test accuracy, essentially matching the 100 % accuracy of a classical ML model on the same task. This indicates that current QML models are at least competitive with classical approaches on small datasets, though not yet vastly superior in raw performance. The theoretical advantage of QML is better seen in certain contrived problems where quantum algorithms show exponential speedup in learning efficiency.
For instance, Huang et al. (2022) proved that a quantum learner could determine properties of a physical system with exponentially fewer experiments than a classical learner in specific scenarios. While such quantum advantage has not been explicitly realized for photocatalysis datasets yet, these proofs-of-concept energize the field by suggesting that as quantum hardware improves, we might tackle currently intractable design spaces (exploring an astronomically large chemical space of catalyst compositions) more feasibly with QML.
Several types of QML models are under investigation for materials design Nandy et al. (2022); Rebentrost et al. (2014); Li et al. (2024). Quantum variational circuits can serve as quantum neural networks: they consist of parameterized quantum gate sequences whose parameters are optimized (via a classical optimizer) to minimize a cost function. These have been used to fit potential energy surfaces and could learn the relationship between atomic structure and photocatalytic activity by directly encoding quantum states of reactants and catalysts. Quantum generative models, like quantum GANs or quantum variational autoencoders, have been theorized as well; a quantum GAN might generate candidate molecular structures in superposition, potentially discovering new catalysts with desired properties using fewer parameters than a classical GAN. Early studies outside photocatalysis have shown that a quantum GAN can model molecular data distributions with high fidelity, hinting that it might recommend novel photocatalyst structures more efficiently.
An area where QML could be particularly impactful is dealing with the quantum nature of excited states and charge dynamics in photocatalysis. Processes like exciton formation, charge separation, and surface reactions are fundamentally quantum-mechanical. Classical ML models often rely on descriptors (band gaps, effective masses) computed via DFT, which might miss subtle many-body effects. QML models, by contrast, can incorporate quantum simulation results more directly or even run quantum subroutines to evaluate properties on the fly. For example, one could envision a QML-driven active learning loop where a quantum computer evaluates the excited-state energy of a proposed catalyst (via a quantum subroutine akin to time-dependent DFT), and the ML model uses that information to decide the next candidate to evaluate. In principle, such a setup could navigate the search space of, say, defect-engineered photocatalysts more efficiently by quantum-accelerating the property predictions inside the loop.
It is important to note that currently QML is largely in the research phase, and clear demonstrations of its superiority on real photocatalysis problems are yet to be seen (Cong et al. 2019). In practice, today’s quantum hardware (NISQ devices) is limited by noise and small numbers of qubits, which restricts the size of materials that can be simulated and the depth of circuits that can be run reliably. Many reported QML successes are on simplified or synthetic tasks (like distinguishing very small molecules or idealized datasets). For instance, one study classified 350 materials into two structural classes using QSVM and classical SVM, both achieved ∼100 % on the simplified task after careful feature selection. Such results show QML can reach classical-level accuracy, but they do not yet prove an advantage.
Researchers often find that without error correction, quantum models may even underperform well-tuned classical models on noisy, complex data. Therefore, a realistic near-term view is that hybrid classical-quantum approaches will be most useful: one might use classical ML for data preprocessing and feature engineering and a quantum model for a specific bottleneck (like computing a quantum kernel or a molecular energy) where it excels.
As quantum devices improve, we anticipate more head-to-head comparisons on practical datasets (using quantum kernels for predicting metal oxide band alignments) to rigorously test whether QML offers speed or accuracy gains. Even if outright performance gains are modest initially, QML provides valuable experience in incorporating quantum data into our ML workflows; for example, using quantum-calculated descriptors (from quantum subroutines) as features in classical models has already enhanced accuracy for small molecular datasets.
QML represents a forward-looking synergy between two cutting-edge fields. For photocatalysis, the ultimate vision is a quantum-infused discovery platform: imagine an algorithm that could evaluate a photocatalyst’s electronic structure on a quantum computer, feed that into an ML model that predicts the catalyst’s performance and then use quantum optimization to suggest how to tweak the catalyst (via doping or structural modifications) to improve efficiency. Achieving this will require overcoming current hardware constraints and developing new QML algorithms tailored to material science problems. The progress so far, though mostly theoretical or demonstrated on model systems, is encouraging. As summarized in Table 7, QML offers theoretical advantages in data efficiency, quantum feature representation, and generative modeling capabilities, even though classical ML remains dominant in current practical applications. Continued advances in quantum computing could make QML a viable component of photocatalyst design pipelines in the coming decade.
Comparison between classical machine learning (ML) and quantum ML in photocatalyst design.
| Aspect | Classical ML | Quantum ML (QML) | Metrics/Outcomes |
|---|---|---|---|
| Algorithm type | SVR, RF, GBR, neural networks (FNN, GAN) predicting band gaps, absorption coefficients, kinetics | Quantum-enhanced models: QSVM, variational quantum circuits, quantum GANs | SVR models on perovskite band gaps: RMSE ≈ 0.36 eV, R2 ≈ 0.965 |
| Feature representation | Hand-crafted elemental and structural descriptors (valence, electronegativity) | Quantum feature maps leveraging Hilbert-space embeddings (QCNN or quantum kernel SVM) | QSVM proof-of-concept shows exponential speed-ups over SVM on synthetic data using HHL algorithm |
| Performance and scaling | ML scales well for large databases; like RF/GBR used for hundreds of thousands of compounds | QML may require fewer training samples and offer theoretical sample complexity advantages | QML demonstrated learning with exponentially fewer experiments |
| Material case study | ML prediction of perovskite labels via SVR and XGBoost: MAE <0.4 eV | Quantum GANs for molecule generation (drug discovery) with tens of parameters outperform classical GANs | Quantum GANs achieved superior performance with fewer parameters |
| Challenges | Needs curated datasets; black-box modeling, interpretability (SHAP, SISSO help) | Hardware noise, small scale, barren-plateau issues, need hybrid architectures | Quantum classifiers often underperform classical unless carefully designed |
| Quantum advantage | Not evident yet in real photocatalyst datasets | Potential for speed-ups in kernel evaluation and quantum sampling; theoretical speed-up in QSVM/HHL | Experimental QSVM and HHL demonstrated potential exponential speed-up |
High-throughput material screening is computationally expensive when using classical DFT calculations. QML can reduce the computational overhead by learning quantum behavior patterns from small datasets and extrapolating them to predict properties of unexplored materials, as exploited by Ajagekar and You (2023) in Figure 13.

The energy-based model utilizes samples from a quantum annealer to map (a) molecular structures to their properties via a (b) GraphConv neural network. (c) The system estimates free energy landscapes and iteratively optimizes candidate molecules with targeted properties using quantum annealing. Reproduced from (Ajagekar and You 2023), npj computational materials, 9, 143 (2023), under the terms of the creative commons attribution 4.0 international license (CC BY 4.0).
6.3 Integration of DFT, AIMD, and ML for photocatalysis
A powerful trend in recent research is the integration of first-principles simulations (like DFT and AIMD) with AI/ML methods to create a virtuous cycle of material optimization. DFT has long been the workhorse for computing properties such as band structures, surface energies, and reaction barriers for photocatalysts. However, DFT calculations can be too slow to screen thousands of candidates or simulate large interface models. ML can act as a surrogate to DFT, learning from a limited set of DFT results and then predicting properties for uncomputed systems almost instantly.
This DFT→ML surrogate approach was used by Huo et al. (2024) to predict band gaps for over 1,200 semiconductor compositions: an SVR model trained on high-throughput DFT results achieved a root-mean-square error of ∼0.36 eV, rapidly pinpointing new compositions in the desired 2–2.5 eV band gap range. Similarly, ML models have been trained to predict formation energies and stability of materials, allowing quick screening of which hypothetical compounds are likely to be synthesizable and robust. An example is the work by Elbaz and Caspary Toroker (2024) on spinel oxides, where a combination of kernel ridge regression and random forest predicted formation energies to within 0.02 eV of DFT values, flagging the most stable compounds for experimental focus.
Beyond static properties, ab initio molecular dynamics (AIMD) simulations provide insight into dynamic processes such as surface reactions, carrier transport, and catalyst stability at finite temperatures. AIMD can capture phenomena like water molecule adsorption and splitting on a catalyst surface or thermal disorder effects on a catalyst’s crystal structure (Cheng and Sprik 2010; Shao et al. 2019). However, AIMD is notoriously expensive, typically limited to <1,000 atoms and <100 ps with DFT accuracy. Here, ML comes to the rescue via machine-learning force fields (MLFF) or interatomic potentials.
By training on a database of DFT-calculated forces and energies (often from many short AIMD trajectories or random structure perturbations), MLFFs (such as neural networks or Gaussian approximation potentials) can emulate DFT accuracy at a tiny fraction of the cost. This enables much longer and larger-scale simulations, for instance, allowing a nanosecond-scale simulation of a photocatalyst nanoparticle interacting with water, which would be infeasible with direct AIMD. Recent works have started to develop MLFFs for photocatalytic materials; for example, a universal graph neural network potential was trained on a diverse set of metal oxides and shown to predict surface interaction energies within ∼0.1 eV of DFT (Ping et al. 2015).
Such potentials make it possible to simulate carrier diffusion lengths and surface recombination probabilities by propagating atomic motions over longer times, thereby connecting directly to properties like charge carrier mobility and lifetime that are critical for photocatalyst efficiency (Guo et al. 2018; Spiegelman et al. 2020). One study reported a constrained AIMD simulation of the water oxidation mechanism on a TiO2 surface (with an organic dye sensitizer), which provided a detailed step-by-step picture of oxygen evolution; the authors noted that extending this to broader time/length scales would require ML-accelerated dynamics due to the prohibitive cost of conventional AIMD (Lee et al. 2007).
Time-dependent DFT (TD-DFT) and many-body techniques (GW/BSE) are also being integrated with ML to tackle excited-state properties like optical absorption spectra and exciton behavior. Photocatalytic performance is governed not only by band gaps but also by exciton binding energies (the energy holding electron-hole pairs together) and charge-carrier effective masses (which influence mobility) (Samanta et al. 2022). Traditional DFT (with semi-local functionals) often misestimates these, so higher-level calculations or experimental data are needed. In a recent large-scale study of organic photocatalysts, TD-DFT was used to compute excitonic energy levels and ionization potentials for hundreds of molecules, which then served as input features for an ML model predicting hydrogen evolution rates (Bonin et al. 2014). The authors specifically calculated the exciton binding energy (difference between electronic and optical band gaps) for each molecule, noting that many had large Eeb that could hinder free charge generation.
By including descriptors like exciton binding energy, singlet–triplet splitting, and dielectric constant, their ML model could recognize which molecules had inherently favorable excited-state dynamics for charge separation. Impressively, the model identified a set of structural motifs (electron-withdrawing groups to reduce exciton binding) that correlated with higher H2 evolution, guiding chemists to focus on those motifs in new designs. This demonstrates the power of marrying advanced electronic-structure calculations with ML: the former contributes physically rich, relevant features, and the latter finds patterns within those features to predict outcomes.
AIMD + ML methods also help evaluate catalyst stability under realistic conditions, which is crucial for practical deployment. For example, AIMD simulations of a photocatalyst in water can reveal if the material’s surface will leach ions or undergo phase transformations. By analyzing many snapshots from AIMD (or using ML to classify trajectories), one can predict the catalyst’s long-term stability or identify which surface facets are prone to deactivation. In one report, AIMD simulations of a TiO2–water interface revealed significant discrepancies in predicted band edge positions when using a standard DFT functional, underscoring the need for careful choice of methods or empirical corrections.
Efforts are underway to use reinforcement learning on AIMD trajectories to discover reaction pathways; for instance, an RL agent can bias the simulation to escape shallow metastable states and find the critical transition state for water splitting on a surface, effectively accelerating rare-event sampling. These hybrid physics-AI approaches constitute a form of “self-driving MD”, where ML guides the simulation to efficiently explore important events (surface bond breaking, defect formation) without brute-force time integration.
Looking ahead, the integration of DFT, AIMD, and ML is expected to become even tighter. We foresee autonomous workflows where: (i) An ML model is trained on an initial batch of DFT data (small set of materials or reactions). (ii) The model predicts a host of new candidates (materials or reaction conditions) likely to perform well. (iii) Those candidates are fed into high-fidelity DFT or AIMD calculations for validation. (iv) The new results are added to the training pool to refine the ML model (active learning loop). Such closed-loop systems can iteratively converge on optimal solutions with far fewer total computations than a brute-force search. Early versions of this paradigm have been demonstrated, for example, to identify promising photocatalytic CO2 reduction materials by cycling between ML predictions and DFT verification. Another future direction is incorporating multi-modal data: experimental characterization (X-ray spectra, microscopy images) could be fused with DFT data in training an ML model, providing a more holistic view of what makes a good photocatalyst. Graph Neural Networks (GNNs), which naturally represent atoms and bonds as graph inputs, are particularly promising for unifying such data and have shown success in learning material properties across chemical and structural diversity.
In conclusion, recent advances illustrate that the synergy of computational physics and machine learning is a powerful accelerator for photocatalyst development. By handling the data deluge from high-throughput computations and experiments, ML extracts actionable knowledge (trends, key descriptors, candidate suggestions) that can be fed back into theory and synthesis. At the same time, anchoring ML models in physical reality through DFT/AIMD data and interpretability techniques ensures that these models do more than just fit the data; they contribute to understanding why certain materials work well.
As the community continues to address challenges like data quality, model generalization, and quantum computing integration, we move closer to a paradigm where discovering a new water-splitting catalyst is less about serendipity and more about design, driven by algorithms that tirelessly learn from every piece of data available. This approach is paving the way toward photocatalysts with the efficiency and stability needed for sustainable hydrogen production, discovered on accelerated timelines unimaginable just a decade ago.
7 DFT–artificial intelligence challenges and future perspective
The integration of Density Functional Theory (DFT) with Artificial Intelligence (AI) has emerged as a powerful approach in accelerating the design and optimization of materials for photocatalytic water splitting. While this synergy offers numerous benefits, such as reducing computational costs and enabling high-throughput screening, it is not without challenges. This section examines the salient challenges associated with the DFT-AI paradigm and explores future perspectives for advancing this integration in photocatalysis research.
7.1 Challenges in DFT–artificial intelligence integration
DFT calculations and experimental data often suffer from inconsistencies, making it difficult to train reliable AI models. Different research groups use varying exchange-correlation functionals, basis sets, and pseudopotentials, leading to non-uniform datasets. Inconsistencies in measurement techniques and environmental conditions can further complicate AI model training. Establish standardized protocols for DFT calculations and experimental validations. Initiatives like the Materials Project and Open Catalyst Project are paving the way by creating large, standardized databases.
While DFT is computationally efficient compared to other quantum mechanical methods, it remains resource-intensive for large-scale or high-throughput studies. Integrating AI models with DFT exacerbates this issue when training on vast datasets or performing iterative calculations. Use surrogate models like Gaussian Process Regression (GPR) or Neural Networks (NNs) to approximate DFT outputs for rapid predictions. Develop hybrid frameworks that combine low-fidelity models with high-fidelity DFT calculations to balance accuracy and efficiency.
AI models, particularly deep learning architectures, often function as black boxes, making it challenging to interpret their predictions and validate results against physical principles. Implement Explainable AI (XAI) techniques, such as SHapley Additive exPlanations (SHAP) or Local Interpretable Model-Agnostic Explanations (LIME), to identify prime features driving predictions. Use physics-informed neural networks (PINNs) to incorporate known physical laws directly into AI models.
AI models trained on specific material systems often struggle to generalize to new compositions or structures, limiting their utility in discovering novel photocatalysts. Employ transfer learning, where pre-trained AI models are fine-tuned for specific material classes with limited data. Use unsupervised learning methods to cluster similar materials and identify transferable patterns.
7.2 Future perspectives in DFT–artificial intelligence integration
Integrating quantum mechanics into AI models represents a significant frontier in improving the accuracy and scalability of DFT-AI systems. Quantum Machine Learning (QML) leverages quantum computing to accelerate the training and inference of AI models, particularly for DFT-like simulations, by processing complex quantum states more efficiently than classical methods. Quantum Neural Networks (QNNs) combine the strengths of quantum systems and AI algorithms, offering powerful tools to solve high-dimensional problems in material science, such as the prediction of electronic structures and reaction dynamics. These quantum-inspired approaches promise to reduce computational costs and unlock novel pathways for designing advanced photocatalysts.
Automated AI-driven workflows are also poised to revolutionize material discovery by enabling inverse design and active learning frameworks. Inverse design uses AI to propose material structures or compositions with desired properties, which are then validated through DFT calculations. Active learning further enhances efficiency by iteratively selecting the most informative data points, optimizing computational and experimental resources. Multi-fidelity modeling provides additional scalability by combining low-cost approximations, such as semi-empirical methods, with high-accuracy DFT calculations. This hybrid approach allows researchers to generate large training datasets for AI models while maintaining precision, bridging the gap between speed and computational intensity.
Finally, expanding data repositories and integrating multi-modal data are essential for advancing DFT-AI integration. Initiatives like the Materials Project, JARVIS-DFT, and Open Catalyst Project provide standardized datasets, but future repositories must also incorporate multi-modal data, such as electronic structures, optical properties, and experimental synthesis conditions. Graph Neural Networks (GNNs) offer a promising solution for representing atomic structures as graphs, enabling efficient learning of material-property relationships. Multi-task learning can further enhance material screening by simultaneously predicting multiple properties, such as bandgap, stability, and defect tolerance. These advancements will facilitate a more comprehensive understanding of photocatalyst performance, accelerating innovation in sustainable energy solutions.
8 Novelty and contributions
This review offers a novel synthesis by triangulating artificial intelligence (AI), density functional theory (DFT), and quantum computing frameworks to analyze and predict performance in binary photocatalytic hydrogen production systems. While previous reviews have focused singularly on experimental materials, DFT simulations, or ML-driven catalyst screening, this work is the first to systematically integrate classical, data-driven, and quantum perspectives. It provides a comparative mapping of software environments (DFT, TDDFT, QML), benchmark parameters (bandgap, conductivity, absorption), and algorithmic frameworks (ensemble ML, hybrid quantum–classical solvers).
Furthermore, the discussion of how variational quantum eigensolvers (VQE), quantum kernel methods, and time-dependent functionals intersect with current challenges in bandgap tuning and electron dynamics modeling provides both theoretical depth and practical guidance. This integrative approach is designed to aid materials scientists, computational chemists, and quantum researchers in developing energy-efficient, scalable catalysts for future hydrogen-based energy systems.
9 Conclusions
The pursuit of sustainable hydrogen production via photocatalytic water splitting has advanced significantly, fueled by innovations in materials science, density functional theory (DFT), and artificial intelligence (AI). This review triangulates these disciplines to provide a comprehensive roadmap for catalyst discovery and evaluation, marking the first integrative study that links AI models, DFT simulations, and quantum computing in the context of binary photocatalytic systems.
Semiconductors like TiO2 and ZnO remain classical benchmarks, yet newer materials such as g-C3N4, BiVO4, and hybrid perovskites like CsPbBr3 exhibit superior solar-to-hydrogen (STH) efficiencies. For instance, BiVO4 demonstrates a band gap of ∼2.4 eV, well-suited for visible light absorption, while g-C3N4 maintains chemical stability with a tunable band gap near 2.7 eV. Tandem systems and upconversion nanoparticles such as NaYF4: Yb,Er have extended the absorption edge below 980 nm, significantly improving photon utilization.
On the computational front, DFT tools like VASP, Quantum ESPRESSO, and SIESTA have enabled the calculation of critical descriptors including band gap, Fermi level, dipole moment, and density of states. For example, DFT-PBE calculations of doped TiO2 systems report reduced band gaps from 3.2 eV to as low as 2.1 eV upon N-doping, thereby enhancing photocatalytic activity. In addition, time-dependent DFT (TDDFT) has been employed to resolve excited-state dynamics and charge separation pathways in materials like ZnIn2S4 and CdS.
AI models, ranging from random forests to graph neural networks, have been deployed for predicting photocatalytic performance with high accuracy (R2 > 0.85), while quantum machine learning (QML) algorithms like Variational Quantum Eigensolvers (VQE) and Quantum Kernel Ridge Regression (QKRR) have begun to approximate ground-state energies and band structures with fewer computational resources. The synergy between DFT and AI is especially impactful for screening multi-component catalysts, where high-throughput workflows can reduce candidate space by >90 % while maintaining discovery accuracy.
Despite these achievements, several obstacles remain. Material degradation under photocatalytic conditions, particularly in perovskites and sulfides, limits long-term deployment. Moreover, many promising photocatalysts fail to exceed 2 % STH efficiency under 1-sun illumination, far from the DOE’s target of 10 %. The lack of curated datasets for model training and the scarcity of quantum-enhanced simulation platforms also limits reproducibility and scalability.
Future directions must prioritize defect and interface engineering to stabilize catalytic performance, particularly in heterostructures like MoS2/g-C3N4, which exhibit synergistic effects in photogenerated charge transfer. Meanwhile, AI-guided robotic synthesis, coupled with real-time spectroscopy, can close the loop for inverse design. Quantum algorithms such as QLS and QAOA hold potential for simulating photoexcited states, with hybrid quantum-classical systems forecasted to become computational workhorses in photocatalytic materials discovery.
This review not only catalogs state-of-the-art strategies for binary photocatalytic hydrogen systems but also charts a unique integrative path forward. The novelty lies in fusing AI, DFT, and quantum technologies to overcome long-standing barriers in catalyst efficiency, stability, and scalability. By fostering cross-disciplinary frameworks and open data infrastructures, the community can accelerate the realization of next-generation photocatalysts, powering the global transition to clean hydrogen energy.
Acknowledgments
The authors gratefully acknowledge the College of Computing at the Georgia Institute of Technology for providing an academic environment and intellectual encouragement that inspired this work. The use of publicly available computational tools, datasets, and literature resources is also sincerely acknowledged. Any opinions, findings, conclusions, or recommendations expressed in this research are solely those of the authors and do not necessarily reflect the views of their respective affiliations.
-
Research ethics: Not applicable.
-
Informed consent: Not applicable.
-
Author contributions: The authors have accepted responsibility for the entire content of this manuscript and approved its submission. D.D.K.W: conceptualization, methodology, validation and visualization, writing – original draft, review and editing. L.G: methodology, validation and visualization, supervision, writing – review and editing. M.D.G: methodology, validation and visualization, supervision, writing – review and editing. All authors reviewed the manuscript.
-
Use of Large Language Models, AI and Machine Learning Tools: Llama 3.2 model was used to support the authors in careful English proofreading. All scientific concepts, analysis, interpretation, and final content were independently conceived, verified, and approved by the author(s). The model did not contribute to the originality, authorship, or intellectual responsibility of the work.
-
Conflict of interest: The authors state no conflict of interest.
-
Research funding: None declared.
-
Data availability: Not applicable.
Nomenclature
- α-method
-
Common DFT or spectroscopy-based method for band gap estimation
- ∆G‡
-
Activation Gibbs free energy
- ∆
-
Energy difference or change
- η
-
Overpotential
- ∆χ
-
Average electronegativity difference
- Eg
-
Band gap energy
- n
-
Principal quantum number
- S
-
Surface recombination velocity
- 2D
-
Two-dimensional
- kT
-
Thermal energy, where k is the Boltzmann
- R
-
Reaction rate (Arrhenius-type dependence)
- A mg−1
-
Amperes per milligram
- AEM
-
Anion exchange membrance
- AFM
-
Atomic force microscopy
- Ag
-
Silver
- AI
-
Artificial intelligence
- AIMD
-
Ab initio molecular dynamics
- ANN
-
Artificial neural network
- ASAP
-
Atomistic simulation advanced platform
- Au
-
Gold
- B
-
Boron
- BiVO4
-
Bismuth vanadate
- BSD
-
Berkeley software/standard distribution
- CASTEP
-
Cambridge serial total energy package
- CB
-
Conduction band
- Cd
-
Cadmium
- CdS
-
Cadmium sulfide
- CdSe
-
Cadmium selenide
- CFD
-
Computational fluid dynamics
- cm−1
-
Wavenumber (spectroscopy)
- CMR
-
Cardiovascular magnetic resonance
- CNN
-
Convolutional neural network
- CNT
-
Carbon nanotube
- Co
-
Cobalt
- Co3O4
-
Cobalt(II,III) oxide
- Co–Pi
-
Cobalt-phosphate catalyst
- CPU
-
Central processing unit
- Cr
-
Chromium
- CUDA
-
Compute unified device architecture
- CuInS2
-
Copper indium sulfide
- DAE
-
Direct air electrolysis
- DDPG
-
Deep deterministic policy gradient
- DFT
-
Density functional theory
- DL
-
Deep learning
- DOE
-
Department of Energy
- e–/h+
-
Electron-hole pair
- ECL
-
Educational community license
- Er3+
-
Erbium
- ET
-
Electron transfer
- ETU
-
Energy transfer upconversion
- eV
-
Electronvolt
- Fe2O3
-
Ferric oxide
- FNNs
-
Feedforward neural networks
- g-C3N4
-
Graphitic carbon nitride
- GAMESS
-
General atomic and molecular electronic structure system
- GANs
-
Generative adversarial networks
- GBDT
-
Gradient boosted decision trees
- GGA
-
Generalized gradient approximation
- GH∗
-
Adsorbed hydrogen intermediate
- GHG
-
Greenhouse gases
- GNN
-
Graph neural network
- GPAW
-
Grid-based projector augmented wave method
- GPL
-
General public license
- GPR
-
Gaussian process regression
- GPU
-
Graphical processing unit
- GW
-
Gigawatt
- GW/BSE
-
Green’s function and Bethe–Salpeter equation approach
- H∗
-
Adsorbed hydrogen atom
- H+/H2
-
Proton/hydrogen couple
- H2O
-
Water molecule
- HER
-
Hydrogen evolution reaction
- HOMO
-
Highest occupied molecular orbital
- HQDs
-
Hybrid quantum dots
- HSE03
-
Heyd-Scuseria-Ernz-Erhof
- ICSD
-
Inorganic crystal structure database
- IQE
-
Internal quantum efficiency
- IrO2
-
Iridium dioxide
- JARVIS-DFT
-
Joint automated repository for various integrated simulations density functional theory; a curated database
- k-fold
-
Cross-validation technique in machine learning
- KNN
-
K-nearest neighbors
- KRR
-
Kernel ridge regression
- KS
-
Kohn-Sham
- LaNiO3
-
Lanthanum nickelate
- LDA
-
Local density approximation
- LEDs
-
Light-emitting diodes
- LGPL
-
GNU lesser general public license
- LightGBM
-
Light gradient-boosting machine
- LIME
-
Local interpretable model-agnostic explanations
- LSPR
-
Localized surface plasmon resonance
- LTH
-
Light-to-hydrogen
- LUMO
-
Lowest unoccupied molecular orbital
- MAE
-
Mean absolute error
- MD
-
Molecular dynamics
- MEC
-
Microbial electrolysis cells
- MIT
-
Massachusetts Institute of Technology
- ML
-
Machine learning
- MLFF
-
Machine-learned force field
- Mn
-
Manganese
- Mo
-
Molybdenum
- mol·m−2 s−1
-
Molar flux per unit area per second
- MoS2
-
Molybdenum disulfide
- MXenes
-
2D transition metal carbides or nitrides
- NaYF4
-
Sodium yttrium fluoride
- Nb
-
Niobium
- NEGF
-
Non-equilibrium Green’s function
- NHE
-
Normal hydrogen electrode
- Ni
-
Nickel
- Ni(OH)2
-
Nickel(II) hydroxide
- Ni2P
-
Nickel phosphide
- Ni3Mo(101)
-
Nickel-molybdenum alloy surface with (101) orientation
- NIR
-
Near-infrared
- nm
-
Nanometre
- NNs
-
Neural networks
- O2
-
Molecular oxygen
- OER
-
Oxygen evolution reaction
- OH∗
-
Adsorbed hydroxyl intermediate
- OOH∗
-
Adsorbed hydroperoxyl intermediate
- Pb
-
Lead
- PBE0
-
Modified Perdew–Burke–Ernzerhof
- PEC
-
Photo-electrochemical catalysis
- PEM
-
Proton exchange membrane
- PINNs
-
Physics-informed neural networks
- Pt
-
Platinum
- PtCu
-
Platinum-copper alloy
- PyCaret
-
Python-based low-code machine learning library
- QE
-
Quantum efficiency
- QML
-
Quantum machine learning
- QNNs
-
Quantum neural networks
- QSVMs
-
Quantum support vector machines
- QY
-
Quantum yield
- RF
-
Random forest
- RFR
-
Random forest regression
- RL
-
Agent reinforcement Learning agent
- RMSE
-
Root mean square error
- RNNs
-
Recurrent neural networks
- Ru
-
Ruthenium
- SAC
-
Single atom catalyst
- SciAI
-
Science and artificial intelligence
- SEM
-
Scanning electron microscope
- SHAP
-
SHapley Additive exPlanations (explainability method)
- SHAP
-
SHapley Additive exPlanations
- Si
-
Silicon
- SIESTA
-
Spanish Initiative for Electronic Simulations with Thousands of Atoms
- SISSO
-
Sure independence screening and sparsifying operation
- STH
-
Solar-to-hydrogen
- SVR
-
Support vector regression
- TDDFT
-
Time-dependent density functional theory
- TEM
-
Transmission electron microscopy
- TiO2
-
Titanium dioxide
- TMD
-
Transition metal dichalcogenide
- UCNPs
-
Upconversion nanoparticles
- UV–vis
-
Ultraviolet–visible spectroscopy
- V
-
Water-splitting potential
- V2DB
-
Valence-to-D-band database
- VAEs
-
Variational autoencoders
- VASP
-
Vienna ab initio simulation package
- VB
-
Valence band
- VBM
-
Valence band maximum
- VQCs
-
Variational quantum circuits
- WO3
-
Tungsten trioxide
- XAI
-
Implement explainable AI
- XC
-
Exchange-correlation
- XGB
-
Extreme gradient boosting
- XPS
-
X-ray photoelectron spectroscopy
- XRD
-
X-ray crystallography
- Yb3+
-
Ytterbium
- ZnO
-
Zinc oxide
- ZnS
-
Zinc sulfide
- ZnTe
-
Zinc telluride
References
Abbass, K., Qasim, M.Z., Song, H., Murshed, M., Mahmood, H., and Younis, I. (2022). A review of the global climate change impacts, adaptation, and sustainable mitigation measures. Environ. Sci. Pollut. Res. 29: 42539–42559, https://doi.org/10.1007/s11356-022-19718-6.Search in Google Scholar PubMed PubMed Central
Abdin, Z., Zafaranloo, A., Rafiee, A., Mérida, W., Lipiński, W., and Khalilpour, K.R. (2020). Hydrogen as an energy vector. Renew. Sustain. Energy Rev. 120: 109620, https://doi.org/10.1016/j.rser.2019.109620.Search in Google Scholar
Abou Zeid, S. and Leprince-Wang, Y. (2024). Advancements in ZnO-based photocatalysts for water treatment: a comprehensive review. Cryst. 14: 611, https://doi.org/10.3390/cryst14070611.Search in Google Scholar
Abraham, B.M., Sinha, P., Halder, P., and Singh, J.K. (2023). Fusing a machine learning strategy with density functional theory to hasten the discovery of 2D MXene-based catalysts for hydrogen generation. J. Mater. Chem. A 11: 8091–8100, https://doi.org/10.1039/D3TA00344B.Search in Google Scholar
Abraham, B.M., Jyothirmai, M.V., Sinha, P., Viñes, F., Singh, J.K., and Illas, F. (2024). Catalysis in the digital age: unlocking the power of data with machine learning. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 14: e1730, https://doi.org/10.1002/wcms.1730.Search in Google Scholar
Agrawal, D., Mahajan, N., Singh, S.A., and Sreedhar, I. (2024). Green hydrogen production pathways for sustainable future with net zero emissions. Fuel 359: 130131, https://doi.org/10.1016/j.fuel.2023.130131.Search in Google Scholar
Ajagekar, A. and You, F. (2023). Molecular design with automated quantum computing-based deep learning and optimization. npj Comput. Mater. 9: 143, https://doi.org/10.1038/s41524-023-01099-0.Search in Google Scholar
Al-Dahidi, S., Hammad, B., Alrbai, M., and Al-Abed, M. (2024). A novel dynamic/adaptive k-nearest neighbor model for the prediction of solar photovoltaic systems’ performance. Results Eng. 22: 102141, https://doi.org/10.1016/j.rineng.2024.102141.Search in Google Scholar
Albero, J., Peng, Y., and García, H. (2020). Photocatalytic CO2 reduction to C2+ products. ACS Catal. 10: 5734–5749, https://doi.org/10.1039/D4TA02838D.Search in Google Scholar
Almeida, R.M.V.S., da Rosa, A.L., Frauenheim, T., and de Almeida, J.S. (2019). Optoelectronic properties of zinc oxide: a first-principles investigation using the Tran–Blaha modified Becke–Johnson potential. Phys. Status Solidi B 256: 1800380, https://doi.org/10.1002/pssb.201800380.Search in Google Scholar
Ansari, A.A. and Sillanpää, M. (2021). Advancement in upconversion nanoparticles based NIR-driven photocatalysts. Renew. Sustain. Energy Rev. 151: 111631, https://doi.org/10.1016/j.rser.2021.111631.Search in Google Scholar
Arabaci, B., Bakir, R., Orak, C., and Yuksel, A. (2025). Predictive modeling of photocatalytic hydrogen production: integrating experimental insights with machine learning on Fe/g-C3N4 catalysts. Ind. Eng. Chem. Res. 64: 5184–5199, https://doi.org/10.1021/acs.iecr.4c03919.Search in Google Scholar
ASAP-SIESTA (2024). Asap-siesta, https://www.simuneatomistics.com/asap/(Accessed 28 Apr 2024).Search in Google Scholar
Astruc, D. (2020). Introduction: nanoparticles in catalysis. Chem. Rev. 120: 461–463, https://doi.org/10.1021/acs.chemrev.8b00696.Search in Google Scholar PubMed
Awad, M. and Khanna, R. (2015). Efficient learning machines: theories, concepts, and applications for engineers and system designers. Apress, Berkeley CA, pp. 268. pp10.1007/978-1-4302-5990-9Search in Google Scholar
Balarabe, B.Y. and Maity, P. (2022). Visible light-driven complete photocatalytic oxidation of organic dye by plasmonic Au-TiO2 nanocatalyst under batch and continuous flow condition. Colloids Surf. A Physicochem. Eng. Asp. 655: 130247, https://doi.org/10.1016/j.colsurfa.2022.130247.Search in Google Scholar
Barik, R.K. and Woods, L.M. (2024). Frictional properties of two-dimensional materials: data-driven machine learning predictive modeling. ACS Appl. Mater. Interfaces 16: 40149–40159, https://doi.org/10.1021/acsami.4c05532.Search in Google Scholar PubMed
Bhattarai, T., Ebong, A., and Raja, M.Y.A. (2024). A review of light-emitting diodes and ultraviolet light-emitting diodes and their applications. Photon. 11: 491, https://doi.org/10.3390/photonics11060491.Search in Google Scholar
Biamonte, J., Wittek, P., Pancotti, N., Rebentrost, P., Wiebe, N., and Lloyd, S. (2017). Quantum machine learning. Nature 549: 195–202, https://doi.org/10.1038/nature23474.Search in Google Scholar PubMed
Bian, H., Li, D., Yan, J., and Liu, S.F. (2021). Perovskite–a wonder catalyst for solar hydrogen production. J. Energy Chem. 57: 325–340, https://doi.org/10.1016/j.jechem.2020.08.057.Search in Google Scholar
Bonin, J., Robert, M., and Routier, M. (2014). Selective and efficient photocatalytic CO2 reduction to CO using visible light and an iron-based homogeneous catalyst. J. Am. Chem. Soc. 136: 16768–16771, https://doi.org/10.1021/ja510290t.Search in Google Scholar PubMed
Breiman, L. (2001). Random forests. Mach. Learn. 45: 5–32, https://doi.org/10.1023/A:1010933404324.10.1023/A:1010933404324Search in Google Scholar
Broxtermann, M., Korte, S., and Jüstel, T. (2017). Mercury-free UV lamp for disinfection and purification of drinking, process, and wastewater: an approach to assessing its innovation potential and possible market entry strategies. J. Bus. Chem. 14: 106, https://doi.org/10.17879/20249613280.Search in Google Scholar
Campos, D., Wayo, D.D.K., De Santis, R.B., Martyushev, D.A., Yaseen, Z.M., Duru, U.I., Saporetti, C.M., and Goliatt, L. (2024). Evolutionary automated radial basis function neural network for multiphase flowing bottom-hole pressure prediction. Fuel 377: 132666, https://doi.org/10.1016/j.fuel.2024.132666.Search in Google Scholar
Cao, N., Xi, M., Li, X., Zheng, J., Qian, L., Dai, Y., Song, X., and Hu, S. (2022). Recent developments in heterogeneous photocatalysts with near-infrared response. Symmetry 14: 2107, https://doi.org/10.3390/sym14102107.Search in Google Scholar
Cerezo, M., Arrasmith, A., Babbush, R., Benjamin, S.C., Endo, S., Fujii, K., McClean, J.R., Mitarai, K., Yuan, X., Cincio, L., et al. (2021). Variational quantum algorithms. Nature Rev. Phys. 3: 625–644, https://doi.org/10.1038/s42254-021-00348-9.Search in Google Scholar
Chanussot, L., Das, A., Goyal, S., Lavril, T., Shuaibi, M., Riviere, M., Tran, K., Heras-Domingo, J., Ho, C., Hu, W., et al. (2021). Open catalyst 2020 (OC20) dataset and community challenges. ACS Catal. 11: 6059–6072, https://doi.org/10.1021/acscatal.0c04525.Search in Google Scholar
Chen, K. and Arnold, F.H. (2020). Engineering new catalytic activities in enzymes. Nature Catal. 3: 203–213, https://doi.org/10.1038/s41929-019-0385-5.Search in Google Scholar
Chen, D., Cheng, Y., Zhou, N., Chen, P., Wang, Y., Li, K., Huo, S., Cheng, P., Peng, P., Zhang, R., et al. (2020). Photocatalytic degradation of organic pollutants using TiO2-based photocatalysts: a review. J. Clean. Prod. 268: 121725, https://doi.org/10.1016/j.jclepro.2020.121725.Search in Google Scholar
Chen, Y., Huang, G., Gao, Y., Chen, Q., and Bi, J. (2022). Up-conversion fluorescent carbon quantum dots decorated covalent triazine frameworks as efficient metal-free photocatalyst for hydrogen evolution. Int. J. Hydrogen Energy 47: 8739–8748, https://doi.org/10.1016/j.ijhydene.2021.12.220.Search in Google Scholar
Cheng, J. and Sprik, M. (2010). Aligning electronic energy levels at the TiO2/H2O interface. Phys. Rev. B Condens. Matter 82: 081406, https://doi.org/10.1103/PhysRevB.82.081406.Search in Google Scholar
Chun, H., Lee, E., Nam, K., Jang, J.H., Kyoung, W., Noh, S.H., and Han, B. (2021). First-principle-data-integrated machine-learning approach for high-throughput searching of ternary electrocatalyst toward oxygen reduction reaction. Chem Catal. 1: 855–869, https://doi.org/10.1016/j.checat.2021.06.001.Search in Google Scholar
Clark, S.J., Segall, M.D., Pickard, C.J., Hasnip, P.J., Probert, M.I.J., Refson, K., and Payne, M.C. (2005). First principles methods using CASTEP. Z. Kristallogr. Cryst. Mater. 220: 567–570, https://doi.org/10.1524/zkri.220.5.567.65075.Search in Google Scholar
Cong, I., Choi, S., and Lukin, M.D. (2019). Quantum convolutional neural networks. Nat. Phys. 15: 1273–1278, https://doi.org/10.1038/s41567-019-0648-8.Search in Google Scholar
Corwin, D.L. (2021). Climate change impacts on soil salinity in agricultural areas. Eur. J. Soil Sci. 72: 842–862, https://doi.org/10.1111/ejss.13010.Search in Google Scholar
Cosmologic-service (2024). Cosmologic-service, http://www.cosmologic-services.de/downloads/TM72-documentation/DOKse23.html (Accessed 24 Apr 2024).Search in Google Scholar
Das, S., Moon, S., Kaur, R., Sharma, G., Kumar, P., and Lavrenčič Stangar, U. (2024). Artificial neural network modeling of photocatalytic degradation of pollutants: a review of photocatalyst, optimum parameters and model topology. Catal. Rev. Sci. Eng. 1: 1–35, https://doi.org/10.1080/01614940.2024.2338131.Search in Google Scholar
Dawood, F., Anda, M., and Shafiullah, G.M. (2020). Hydrogen production for energy: an overview. Int. J. Hydrogen Energy 45: 3847–3869, https://doi.org/10.1016/j.ijhydene.2019.12.059.Search in Google Scholar
Deringer, V.L., Bartók, A.P., Bernstein, N., Wilkins, D.M., Ceriotti, M., and Csányi, G. (2021). Gaussian process regression for materials and molecules. Chem. Rev. 121: 10073–10141, https://doi.org/10.1021/acs.chemrev.1c00022.Search in Google Scholar PubMed PubMed Central
Di Liberto, G., Tosoni, S., and Pacchioni, G. (2019). Theoretical treatment of semiconductor heterojunctions for photocatalysis: the WO3/BiVO4 interface. J. Phys.: Condens. Matter 31: 434001, https://doi.org/10.1088/1361-648X/ab2fa4.Search in Google Scholar PubMed
Doughan, S., Uddayasankar, U., and Krull, U.J. (2015). A paper-based resonance energy transfer nucleic acid hybridization assay using upconversion nanoparticles as donors and quantum dots as acceptors. Anal. Chim. Acta 878: 1–8, https://doi.org/10.1016/j.aca.2015.04.036.Search in Google Scholar PubMed
Elbaz, Y. and Caspary Toroker, M. (2024). From density functional theory to machine learning predictive models for electrical properties of spinel oxides. Sci. Rep. 14: 12150, https://doi.org/10.1038/s41598-024-62788-4.Search in Google Scholar PubMed PubMed Central
Erhart, P., Albe, K., and Klein, A. (2006). First-principles study of intrinsic point defects in ZnO: role of band structure, volume relaxation, and finite-size effects. Phys. Rev. B Condens. Matter 73: 205203, https://doi.org/10.1103/PhysRevB.73.205203.Search in Google Scholar
Gao, X., Yang, N., Feng, J., Liao, J., Hou, S., Ma, X., Su, D., Yu, X., Yang, Z., Safaei, J., et al. (2023). Defect and interface control on graphitic carbon nitrides/upconversion nanocrystals for enhanced solar hydrogen production. Natl. Sci. Open 2: 20220037, https://doi.org/10.1360/nso/20220037.Search in Google Scholar
Genovese, L., Ospici, M., Deutsch, T., Méhaut, J.F., Neelov, A., and Goedecker, S. (2009). Density functional theory calculation on many-cores hybrid central processing unit–graphic processing unit architectures. J. Chem. Phys. 131: 034103, https://doi.org/10.1063/1.3176496.Search in Google Scholar
Genovese, L., Videau, B., Deutsch, T., Tran, H., and Goedecker, S. (2011). Improvements of BigDFT code in modern HPC architectures. Partnersh. Adv. Comput. Europ., https://prace-ri.eu/wp-content/uploads/Improvements_of_BigDFT_code_in_modern_HPC_architectures.pdf (Accessed 31 Jul 2025).Search in Google Scholar
Giannozzi, P., Baroni, S., Bonini, N., Calandra, M., Car, R., Cavazzoni, C., Ceresoli, D., Chiarotti, G.L., Cococcioni, M., Dabo, I., et al. (2009). Quantum ESPRESSO: a modular and open-source software project for quantum simulations of materials. J. Phys.: Condens. Matter 21: 395502, https://doi.org/10.1088/0953-8984/21/39/395502.Search in Google Scholar PubMed
GitHub (2024). Moleculardynamics-repository, https://github.com/search?q=molecular%20dynamicandtype=repositories (Accessed 25 Apr 2024).Search in Google Scholar
Gladkikh, V., Kim, D.Y., Hajibabaei, A., Jana, A., Myung, C.W., and Kim, K.S. (2020). Machine learning for predicting the band gaps of ABX3 perovskites from elemental properties. J. Phys. Chem. C 124: 8905–8918, https://doi.org/10.1021/acs.jpcc.9b11768.Search in Google Scholar
Glorot, X. and Bengio, Y. (2010). Understanding the difficulty of training deep feedforward neural networks. Proceedings of the Thirteenth International Conference on Artificial Intelligence and Statistics (AISTATS 2010). JMLR Workshop and Conference Proceedings 9: 249–256, https://proceedings.mlr.press/v9/glorot10a (Accessed 31 Jul 2025).Search in Google Scholar
Gomes, S.A.O., Vieira, C.S., Almeida, D.B., Santos-Mallet, J.R., Menna-Barreto, R.F.S., Cesar, C.L., and Feder, D. (2011). CdTe and CdSe quantum dots cytotoxicity: a comparative study on microorganisms. Sens. 11: 11664–11678, https://doi.org/10.3390/s111211664.Search in Google Scholar PubMed PubMed Central
Gonze, X., Amadon, B., Anglade, P.M., Beuken, J.M., Bottin, F., Boulanger, P., Bruneval, F., Caliste, D., Caracas, R., Côté, M., et al. (2009). ABINIT: first-principles approach to material and nanosystem properties. Comput. Phys. Commun. 180: 2582–2615, https://doi.org/10.1016/j.cpc.2009.07.007.Search in Google Scholar
Goodarzi, N., Ashrafi-Peyman, Z., Khani, E., and Moshfegh, A.Z. (2023). Recent progress on semiconductor heterogeneous photocatalysts in clean energy production and environmental remediation. Catal. 13: 1102, https://doi.org/10.3390/catal13071102.Search in Google Scholar
Gui, X., Lu, Y., Wang, Q., Cai, M., and Sun, S. (2024). Application of quantum dots for photocatalytic hydrogen evolution reaction. Appl. Sci. 14: 5333, https://doi.org/10.3390/app14125333.Search in Google Scholar
Guo, N., Yam, K.M., and Zhang, C. (2018). Substrate engineering of graphene reactivity: towards high-performance graphene-based catalysts. npj 2D Mater. Appl. 2: 1, https://doi.org/10.1038/s41699-017-0046-y.Search in Google Scholar
Hafner, J. (2008). Ab-initio simulations of materials using VASP: density-functional theory and beyond. J. Comput. Chem. 29: 2044–2078, https://doi.org/10.1002/jcc.21057.Search in Google Scholar PubMed
Haghshenas, Y., Wong, W.P., Gunawan, D., Khataee, A., Keyikoğlu, R., Razmjou, A., Kumar, P.V., Toe, C.Y., Masood, H., Amal, R., et al. (2024). Predicting the rates of photocatalytic hydrogen evolution over cocatalyst-deposited TiO2 using machine learning with active photon flux as a unifying feature. EES Catal. 2: 612–623, https://doi.org/10.1039/D3EY00246B.Search in Google Scholar
Hammer, B., Hansen, L.B., and Nørskov, J.K. (1999). Improved adsorption energetics within density-functional theory using revised Perdew–Burke–Ernzerhof functionals. Phys. Rev. B 59: 7413, https://doi.org/10.1103/PhysRevB.59.7413.Search in Google Scholar
Horikoshi, S., Yamamoto, D., Hagiwara, K., Tsuchida, A., Matsumoto, I., Nishiura, Y., Kiyoshima, Y., and Serpone, N. (2019). Development of a Hg-free UV light source and its performance in photolytic and photocatalytic applications. Photochem. Photobiol. Sci. 18: 328–335, https://doi.org/10.1039/c8pp00395e.Search in Google Scholar PubMed
Hu, R., Lei, W., Yuan, H., Han, S., and Liu, H. (2022). High-throughput prediction of the band gaps of van der Waals heterostructures via machine learning. Nanomater. 12: 2301, https://doi.org/10.3390/nano12132301.Search in Google Scholar PubMed PubMed Central
Huang, H.Y., Broughton, M., Mohseni, M., Babbush, R., Boixo, S., Neven, H., and McClean, J.R. (2021). Power of data in quantum machine learning. Nat. Commun. 12: 2631, https://doi.org/10.1038/s41467-021-22539-9.Search in Google Scholar PubMed PubMed Central
Huang, H.Y., Broughton, M., Cotler, J., Chen, S., Li, J., Mohseni, M., Neven, H., Babbush, R., Kueng, R., Preskill, J., et al. (2022). Quantum advantage in learning from experiments. Sci. 376: 1182–1186, https://doi.org/10.1126/science.abn729.Search in Google Scholar
Huang, M., Yu, M., Si, R., Zhao, X., Chen, S., Liu, K., and Pan, X. (2023). Tailoring morphology in hydrothermally synthesized CdS/ZnS nanocomposites for extraordinary photocatalytic H2 generation via type-II heterojunction. Catal. 13: 1123, https://doi.org/10.3390/catal13071123.Search in Google Scholar
Huo, S., Zhang, S., Wu, Q., and Zhang, X. (2024). Feature-assisted machine learning for predicting band gaps of binary semiconductors. Nanomater. 14: 445, https://doi.org/10.3390/nano14050445.Search in Google Scholar PubMed PubMed Central
Irawan, S., Wayo, D.D.K., Bayramov, E., Fathaddin, M.T., and Goliatt, L. (2024). Molecular dynamic prognosis for Ti-C10H16N2O8 filter cake decomposition. In: SPE annual caspian technical conference. Society of Petroleum Engineers (SPE), p. D021S007R002 https://onepetro.org/SPECTCE/proceedings-abstract/24CTC/24CTC/607248 (Accessed 31 Jul 2025).10.2118/223416-MSSearch in Google Scholar
Isazawa, T. and Cole, J.M. (2023). Automated construction of a photocatalysis dataset for water-splitting applications. Sci. Data 10: 651, https://doi.org/10.1038/s41597-023-02511-6.Search in Google Scholar PubMed PubMed Central
Jacak, L., Hawrylak, P., and Wojs, A. (2013). Quantum dots. Springer Science+Business Media, https://books.google.kz/books?hl=en&lr=&id=wW_tCAAAQBAJ (Accessed 31 Jul 2025).Search in Google Scholar
Javed, M.F., Shahab, M.Z., Asif, U., Najeh, T., Aslam, F., Ali, M., and Khan, I. (2024). Evaluation of machine learning models for predicting TiO2 photocatalytic degradation of air contaminants. Sci. Rep. 14: 13688, https://doi.org/10.1038/s41598-024-64486-7.Search in Google Scholar PubMed PubMed Central
Jayan, K.D. and Babu, K. (2024). Luminescent perovskite quantum dots: progress in fabrication, modelling and machine learning approaches for advanced photonic and quantum computing applications. J. Lumin. 268: 120906, https://doi.org/10.1016/j.jlumin.2024.120906.Search in Google Scholar
Jiang, D., Sun, Z., Jia, H., Lu, D., and Du, P. (2016). A cocatalyst-free CdS nanorod/ZnS nanoparticle composite for high-performance visible-light-driven hydrogen production from water. J. Mater. Chem. A 4: 675–683, https://doi.org/10.1039/C5TA07420G.Search in Google Scholar
Jin, B., Cho, Y., Park, C., Jeong, J., Kim, S., Jin, J., Kim, W., Wang, L., Lu, S., Zhang, S., et al. (2022). A two-photon tandem black phosphorus quantum dot-sensitized BiVO4 photoanode for solar water splitting. Energy Environ. Sci. 15: 672–679, https://doi.org/10.1039/D1EE03014K.Search in Google Scholar
Jyothirmai, M.V., Roshini, D., Abraham, B.M., and Singh, J.K. (2023). Accelerating the discovery of g-C3N4-supported single atom catalysts for hydrogen evolution reaction: a combined DFT and machine learning strategy. ACS Appl. Energy Mater. 6: 5598–5606, https://doi.org/10.1021/acsaem.3c00835.Search in Google Scholar
Jyothirmai, M.V., Dantuluri, R., Sinha, P., Abraham, B.M., and Singh, J.K. (2024). Machine-learning-driven high-throughput screening of transition-metal atom intercalated g-C3N4/MX2 (M = Mo, W; X = S, Se, Te) heterostructures for the hydrogen evolution reaction. ACS Appl. Mater. Interfaces 16: 12437–12445, https://doi.org/10.1021/acsami.3c17389.Search in Google Scholar PubMed
Kao, P.Y., Yang, Y.C., Chiang, W.Y., Hsiao, J.Y., Cao, Y., Aliper, A., Ren, F., Aspuru-Guzik, A., Zhavoronkov, A., Hsieh, M.H., et al. (2023). Exploring the advantages of quantum generative adversarial networks in generative chemistry. J. Chem. Inf. Model. 63: 3307–3318, https://doi.org/10.1021/acs.jcim.3c00562.Search in Google Scholar PubMed PubMed Central
Karimi-Maleh, H., Kumar, B.G., Rajendran, S., Qin, J., Vadivel, S., Durgalakshmi, D., Gracia, F., Soto-Moscoso, M., Orooji, Y., and Karimi, F. (2020). Tuning of metal oxides photocatalytic performance using Ag nanoparticles integration. J. Mol. Liq. 314: 113588, https://doi.org/10.1016/j.molliq.2020.113588.Search in Google Scholar
Karl, T.R., Williams Jr, C.N., Young, P.J., and Wendland, W.M. (2009). Global climate change impacts in the United States: a state of knowledge report. Cambridge University Press, New York, https://books.google.kz/books?hl=en&lr=&id=UCg7inA-HksC (Accessed 31 Jul 2025).Search in Google Scholar
Khodadadian, M. (2019). Optimizing photon utilization in LED-based photocatalytic reactors, PhD. thesis. Delft University of Technology, Delft.Search in Google Scholar
Ko, M., Kim, Y., Woo, J., Lee, B., Mehrotra, R., Sharma, P., Kim, J., Hwang, S.W., Jeong, H.Y., Lim, H., et al. (2022). Direct propylene epoxidation with oxygen using a photo-electro-heterogeneous catalytic system. Nature Catal. 5: 37–44, https://doi.org/10.1038/s41929-021-00724-9.Search in Google Scholar
Kohn, W. and Sham, L.J. (1965). Self-consistent equations including exchange and correlation effects. Phys. Rev. 140: A1133, https://doi.org/10.1103/PhysRev.140.A1133.Search in Google Scholar
Kowsari, K., Jafari Meimandi, K., Heidarysafa, M., Mendu, S., Barnes, L., and Brown, D. (2019). Text classification algorithms: a survey. Information 10: 150, https://doi.org/10.3390/info10040150.Search in Google Scholar
Kubiak, A. (2024). Impact of LED radiation intensity on gold nanoparticles photodeposition on TiO2 with physicochemical and photocatalytic characterization. Sci. Rep. 14: 20563, https://doi.org/10.1038/s41598-024-71605-x.Search in Google Scholar PubMed PubMed Central
Kumar, R. and Singh, A.K. (2021). Chemical hardness-driven interpretable machine learning approach for rapid search of photocatalysts. npj Comput. Mater. 7: 197, https://doi.org/10.1038/s41524-021-00669-4.Search in Google Scholar
Kumar, M., Singh, N.K., Kumar, R.S., and Singh, R. (2024). Production of green hydrogen through photocatalysis. In: Towards sustainable and green hydrogen production by photocatalysis: insights into design and development of efficient materials, 2. ACS Publications, Washington, DC, pp. 1–24.10.1021/bk-2024-1468.ch001Search in Google Scholar
Lee, K.S. and El-Sayed, M.A. (2006). Gold and silver nanoparticles in sensing and imaging: sensitivity of plasmon response to size, shape, and metal composition. J. Phys. Chem. B 110: 19220–19225, https://doi.org/10.1021/jp062536y.Search in Google Scholar PubMed
Lee, Y., Terashima, H., Shimodaira, Y., Teramura, K., Hara, M., Kobayashi, H., Domen, K., and Yashima, M. (2007). Zinc germanium oxynitride as a photocatalyst for overall water splitting under visible light. J. Phys. Chem. C 111: 1042–1048, https://doi.org/10.1021/jp0656532.Search in Google Scholar
Lee, W.Y., Lee, J., Kwon, H.J., Kim, K., Kang, H., and Jang, J. (2024). High-detectivity silver telluride nanoparticle-based near-infrared photodetectors functionalized with surface-plasmonic gold nanoparticles. Appl. Surf. Sci. 654: 159563, https://doi.org/10.1016/j.apsusc.2024.159563.Search in Google Scholar
Leimkuhler, B. and Matthews, C. (2015). Molecular dynamics: with deterministic and stochastic numerical methods. In: Antman, S.S., Holmes, P., and Greengard, L. (Eds.). Interdisciplinary Applied Mathematics (IAM), Vol. XXXIX. Springer, Berlin, Heidelberg.10.1007/978-3-319-16375-8Search in Google Scholar
Leng, M., Chen, Z., Yang, Y., Li, Z., Zeng, K., Li, K., Niu, G., He, Y., Zhou, Q., and Tang, J. (2016). Lead-free, blue emitting bismuth halide perovskite quantum dots. Angew. Chem., Int. Ed. 55: 15012–15016, https://doi.org/10.1002/anie.201608160.Search in Google Scholar PubMed
Li, W., Jacobs, R., and Morgan, D. (2018). Predicting the thermodynamic stability of perovskite oxides using machine learning models. Comput. Mater. Sci. 150: 454–463, https://doi.org/10.1016/j.commatsci.2018.04.033.Search in Google Scholar
Li, X., Maffettone, P.M., Che, Y., Liu, T., Chen, L., and Cooper, A.I. (2021). Combining machine learning and high-throughput experimentation to discover photocatalytically active organic molecules. Chem. Sci. 12: 10742–10754, https://doi.org/10.1039/D1SC02150H.Search in Google Scholar PubMed PubMed Central
Li, C., Liu, J., Li, H., Wu, K., Wang, J., and Yang, Q. (2022a). Covalent organic frameworks with high quantum efficiency in sacrificial photocatalytic hydrogen evolution. Nat. Commun. 13: 2357, https://doi.org/10.1038/s41467-022-30035-x.Search in Google Scholar PubMed PubMed Central
Li, Z., Yoon, J., Zhang, R., Rajabipour, F., Srubar III, W.V., Dabo, I., and Radlinska, A. (2022b). Machine learning in concrete science: applications, challenges, and best practices. npj Comput. Mater. 8: 127, https://doi.org/10.1038/s41524-022-00810-x.Search in Google Scholar
Li, J., Li, Y., Song, J., Zhang, J., and Zhang, S. (2024). Quantum support vector machine for classifying noisy data. IEEE Trans. Comput. 73: 2233–2247, https://doi.org/10.1109/TC.2024.3416619.Search in Google Scholar
Li, J., Wang, J., He, T., Li, Z., Xu, H., Fan, Z., Liao, F., Liu, Y., and Kang, Z. (2025a). Dynamic machine learning-driven optimization of microwave-synthesized photocatalysts for enhanced hydrogen peroxide production. ChemCatChem 17: e00341, https://doi.org/10.1002/cctc.202500341.Search in Google Scholar
Li, Y., Yu, J., Li, J., Wang, Y., and Sun, B. (2025b). Critical factors and equilibrium analysis of luminescent down-shifting process for silicon heterojunction solar cells. Adv. Energy Mater. 15: 2405918, https://doi.org/10.1002/aenm.202405918.Search in Google Scholar
Lim, P.F., Leong, K.H., Sim, L.C., Saravanan, P., and Aziz, A.A. (2019). Perovskite oxide–based photocatalysts for excellent visible light–driven photocatalysis and energy conversion. In: Tripathi, D.K., Ahmad, P., Sharma, S., and Chauhan, D.K. (Eds.). Plant nanobionics. Volume 2. Approaches in nanoparticles, biosynthesis, and toxicity. Springer, Cham, pp. 35–54.10.1007/978-3-030-16379-2_2Search in Google Scholar
Link, S., Wang, Z.L., and El-Sayed, M.A. (1999). Alloy formation of gold-silver nanoparticles and the dependence of the plasmon absorption on their composition. J. Phys. Chem. B 103: 3529–3533, https://doi.org/10.1021/jp990387w.Search in Google Scholar
Liu, Y., Xu, Y., Geng, X., Huo, Y., Chen, D., Sun, K., Zhou, G., Chen, B., and Tao, K. (2018). Synergistic targeting and efficient photodynamic therapy based on graphene oxide quantum dot-upconversion nanocrystal hybrid nanoparticles. Small 14: 1800293, https://doi.org/10.1002/smll.201800293.Search in Google Scholar PubMed
Liu, Q., Pan, K., Lu, Y., Wei, W., Wang, S., Du, W., Ding, Z., and Zhou, Y. (2022). Data-driven for accelerated design strategy of photocatalytic degradation activity prediction of doped TiO2 photocatalyst. J. Water Proc. Eng. 49: 103126, https://doi.org/10.1016/j.jwpe.2022.103126.Search in Google Scholar
Lu, B., Xia, Y., Ren, Y., Xie, M., Zhou, L., Vinai, G., Morton, S.A., Wee, A.T.S., van der Wiel, W.G., Zhang, W., et al. (2024). When machine learning meets 2D materials: a review. Adv. Sci. 11: 2305277, https://doi.org/10.1002/advs.202305277.Search in Google Scholar PubMed PubMed Central
Ma, Y., Lin, L., Takata, T., Hisatomi, T., and Domen, K. (2023). A perspective on two pathways of photocatalytic water splitting and their practical application systems. Phys. Chem. Chem. Phys. 25: 6586–6601, https://doi.org/10.1039/D2CP05427B.Search in Google Scholar PubMed
Masood, H., Toe, C.Y., Teoh, W.Y., Sethu, V., and Amal, R. (2019). Machine learning for accelerated discovery of solar photocatalysts. ACS Catal. 9: 11774–11787, https://doi.org/10.1021/acscatal.9b02531.Search in Google Scholar
Masood, H., Sirojan, T., Toe, C.Y., Kumar, P.V., Haghshenas, Y., Sit, P.H.L., Amal, R., Sethu, V., and Teoh, W.Y. (2023). Enhancing prediction accuracy of physical band gaps in semiconductor materials. Cell Rep. Phys. Sci. 4: 101555, https://doi.org/10.1016/j.xcrp.2023.101555.Search in Google Scholar
mat3ra (2024). mat3ra, https://www.mat3ra.com/(Accessed 28 Apr 2024).Search in Google Scholar
Medsker, L.R. and Jain, L.C. (2001). Recurrent neural networks: design and applications. CRC Press, Boca Raton.Search in Google Scholar
Meinhardová, V., Dubnová, L., Drobná, H., Matějová, L., Kočí, K., and Čapek, L. (2023). Role of lamp type in conventional batch and micro-photoreactor for photocatalytic hydrogen production. Front. Chem. 11: 1271410, https://doi.org/10.3389/fchem.2023.1271410.Search in Google Scholar PubMed PubMed Central
Meyer, B.K., Polity, A., Reppin, D., Becker, M., Hering, P., Klar, P.J., Sander, T., Reindl, C., Benz, J., Eickhoff, M., et al. (2012). Binary copper oxide semiconductors: from materials towards devices. Phys. Status Solidi B 249: 1487–1509, https://doi.org/10.1002/pssb.201248128.Search in Google Scholar
Mienye, I.D., Swart, T.G., and Obaido, G. (2024). Recurrent neural networks: a comprehensive review of architectures, variants, and applications. Information 15: 517, https://doi.org/10.3390/info15090517.Search in Google Scholar
Mihandoost, S., Rezvantalab, S., Pallares, R.M., Schulz, V., and Kiessling, F. (2024). A generative adversarial network approach to predict nanoparticle size in microfluidics. ACS Biomater. Sci. Eng. 11: 268–279, https://doi.org/10.1021/acsbiomaterials.4c01423.Search in Google Scholar PubMed
Miyoshi, A., Nishioka, S., and Maeda, K. (2018). Water splitting on rutile TiO2-based photocatalysts. Chem. Eur J. 24: 18204–18219, https://doi.org/10.1002/chem.201800799.Search in Google Scholar PubMed
Nandy, A., Duan, C., and Kulik, H.J. (2022). Audacity of huge: overcoming challenges of data scarcity and data quality for machine learning in computational materials discovery. Curr. Opin. Chem. Eng. 36: 100778, https://doi.org/10.1016/j.coche.2021.100778.Search in Google Scholar
Ndia Ntone, E.P., Abu Samah, R., Abdul Wahab, M.S., Alsalhy, Q.F., and Rahman, S.A. (2024). Review of PPCPs remediation in Asia: the role of agricultural waste in adsorption-membrane hybrid technology. Chem. Eng. Commun. 212: 1132–1166, https://doi.org/10.1080/00986445.2024.2447843.Search in Google Scholar
Neese, F. (2012). The ORCA program system. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2: 73–78, https://doi.org/10.1002/wcms.81.Search in Google Scholar
Nisar, M., Khan, N., Qadir, M.I., and Shah, Z. (2025). Recent advances in TiO2-based photocatalysts for efficient water splitting to hydrogen. Nanomater. 15: 984, https://doi.org/10.3390/nano15130984.Search in Google Scholar PubMed PubMed Central
Odabasi Lee, S., Lakhera, S.K., and Yong, K. (2023). Strategies to enhance interfacial spatial charge separation for high-efficiency photocatalytic overall water-splitting: a review. Adv. Energy Sustain. Res. 4: 2300130, https://doi.org/10.1002/aesr.202300130.Search in Google Scholar
Oh, S.H.V., Yoo, S.H., and Jang, W. (2024). Small dataset machine-learning approach for efficient design space exploration: engineering ZnTe-based high-entropy alloys for water splitting. npj Comput. Mater. 10: 166, https://doi.org/10.1038/s41524-024-01341-3.Search in Google Scholar
Orio, M., Pantazis, D.A., and Neese, F. (2009). Density functional theory. Photosynth. Res. 102: 443–453, https://doi.org/10.1007/s11120-009-9404-8.Search in Google Scholar PubMed PubMed Central
Pakornchote, T., Choomphon-Anomakhun, N., Arrerut, S., Atthapak, C., Khamkaeo, S., Chotibut, T., and Bovornratanaraks, T. (2024). Diffusion probabilistic models enhance variational autoencoder for crystal structure generative modeling. Sci. Rep. 14: 1275, https://doi.org/10.1038/s41598-024-51400-4.Search in Google Scholar PubMed PubMed Central
Patel, V., Sani, R.K., Sharma, A.K., and Bagri, U. (2024). Hydrogen production scale-up. In: Pandey, A., Mohan, S.V., Khanal, S.K., and Biswas, B. (Eds.). Renewable hydrogen. Elsevier, Amsterdam, pp. 83–104.10.1016/B978-0-323-95379-5.00002-XSearch in Google Scholar
Perdew, J.P. (1985). Density functional theory and the band gap problem. Int. J. Quantum Chem. 28: 497–523, https://doi.org/10.1002/qua.560280846.Search in Google Scholar
Perdew, J.P. (1986). Density-functional approximation for the correlation energy of the inhomogeneous electron gas. Phys. Rev. B 33: 8822, https://doi.org/10.1103/PhysRevB.33.8822.Search in Google Scholar
Perdew, J.P., Chevary, J.A., Vosko, S.H., Jackson, K.A., Pederson, M.R., Singh, D.J., and Fiolhais, C. (1992). Atoms, molecules, solids, and surfaces: applications of the generalized gradient approximation for exchange and correlation. Phys. Rev. B 46: 6671, https://doi.org/10.1103/PhysRevB.46.6671.Search in Google Scholar PubMed
Perdew, J.P., Burke, K., and Ernzerhof, M. (1996). Generalized gradient approximation made simple. Phys. Rev. Lett. 77: 3865, https://doi.org/10.1103/PhysRevLett.77.3865.Search in Google Scholar PubMed
Pereira, F., Xiao, K., Latino, D.A.R.S., Wu, C., Zhang, Q., and Aires-de Sousa, J. (2017). Machine learning methods to predict density functional theory B3LYP energies of HOMO and LUMO orbitals. J. Chem. Inf. Model. 57: 11–21, https://doi.org/10.1021/acs.jcim.6b00340.Search in Google Scholar PubMed
Peterson, L.E. (2009). K-nearest neighbor. Scholarpedia 4: 1883, https://doi.org/10.4249/scholarpedia.1883, http://scholarpedia.org/article/K-nearest_neighbor (Accessed 1 Aug 2025).Search in Google Scholar
Pham, N.N.T., Kang, S.G., Kim, H.J., Pak, C., Han, B., and Lee, S.G. (2021). Catalytic activity of Ni3Mo surfaces for hydrogen evolution reaction: a density functional theory approach. Appl. Surf. Sci. 537: 147894, https://doi.org/10.1016/j.apsusc.2020.147894.Search in Google Scholar
Pihosh, Y., Turkevych, I., Mawatari, K., Uemura, J., Kazoe, Y., Kosar, S., Makita, K., Sugaya, T., Matsui, T., Fujita, D., et al. (2015). Photocatalytic generation of hydrogen by core–shell WO3/BiVO4 nanorods with ultimate water splitting efficiency. Sci. Rep. 5: 11141, https://doi.org/10.1038/srep11141.Search in Google Scholar PubMed PubMed Central
Ping, Y., Sundararaman, R., and Goddard, W.A. (2015). Solvation effects on the band edge positions of photocatalysts from first principles. Phys. Chem. Chem. Phys. 17: 30499–30509, https://doi.org/10.1039/C5CP05740J.Search in Google Scholar PubMed
Priya, R.L., Hariprasad, B.S., Dhayanithi, C.A., Paunkumar, P., Bhuvaneswari, C., and Babu, S.G. (2024). Photocatalytic and photoelectrochemical technologies for hydrogen production: commercialization aspect. In: Photocatalysis for energy and environmental applications: current trends and future perspectives. Springer, Singapore, pp. 299–325.10.1007/978-981-97-1939-6_10Search in Google Scholar
Qi, Y., Zhang, J., Kong, Y., Zhao, Y., Chen, S., Li, D., Liu, W., Chen, Y., Xie, T., Cui, J., et al. (2022). Unraveling of cocatalysts photodeposited selectively on facets of BiVO4 to boost solar water splitting. Nat. Commun. 13: 484, https://doi.org/10.1038/s41467-022-28146-6.Search in Google Scholar PubMed PubMed Central
Qi, C., Guo, X., Lu, B., Ruan, B., and Li, P. (2023). Enhanced photocatalytic performance of a ZnO/CdS heterostructure for hydrogen production and mixed dye degradation. ChemistrySelect 8: e202203227, https://doi.org/10.1002/slct.202203227.Search in Google Scholar
Qureshi, A., Wahab, M.A., Badreldin, A., Abdel-Wahab, A., Castaneda, H., and Abdala, A. (2024). Advances in photothermal catalysts for solar-driven hydrogen production. Int. J. Hydrogen Energy 96: 160–181, https://doi.org/10.1016/j.ijhydene.2024.11.124.Search in Google Scholar
Rebentrost, P., Mohseni, M., and Lloyd, S. (2014). Quantum support vector machine for big data classification. Phys. Rev. Lett. 113: 130503, https://doi.org/10.1103/PhysRevLett.113.130503.Search in Google Scholar PubMed
Ren, T., Wang, L., Chang, C., and Li, X. (2020). Machine learning-assisted multiphysics coupling performance optimization in a photocatalytic hydrogen production system. Energy Convers. Manage. 216: 112935, https://doi.org/10.1016/j.enconman.2020.112935.Search in Google Scholar
Rigatti, S.J. (2017). Random forest. Journal of Insurance Medicine 47: 31–39, https://doi.org/10.17849/insm-47-01-31-39.1.Search in Google Scholar PubMed
Sabagh Moeini, A., Shariatmadar Tehrani, F., and Naeimi-Sadigh, A. (2024). Machine learning-enhanced band gaps prediction for low-symmetry double and layered perovskites. Sci. Rep. 14: 26736, https://doi.org/10.1038/s41598-024-77081-7.Search in Google Scholar PubMed PubMed Central
Samanta, B., Morales-Garcia, A., Illas, F., Goga, N., Anta, J.A., Calero, S., Bieberle-Hutter, A., Libisch, F., Muñoz-Garcia, A.B., Pavone, M., et al. (2022). Challenges of modeling nanostructured materials for photocatalytic water splitting. Chem. Soc. Rev. 51: 3794–3818, https://doi.org/10.1039/D1CS00648G.Search in Google Scholar
Sarina, S., Waclawik, E.R., and Zhu, H. (2013). Photocatalysis on supported gold and silver nanoparticles under ultraviolet and visible light irradiation. Green Chem. 15: 1814–1833, https://doi.org/10.1039/C3GC40450A.Search in Google Scholar
Sakar, M., Mithun Prakash, R., and Do, T.O. (2019). Insights into the TiO2-based photocatalytic systems and their mechanisms. Catal. 9: 680, https://doi.org/10.3390/catal9080680.Search in Google Scholar
Sarkar, A., Sharma, S., Balarabe, B.Y., Vadivel, S., Tyagi, T., Das, D., Kumar, S., Puzari, A., and Paul, B. (2024). In-situ synthesis of star-shaped Ni supported on WO3 nanoparticles for selective hydrogenation of cinnamaldehyde. Int. J. Hydrogen Energy 59: 1174–1182, https://doi.org/10.1016/j.ijhydene.2024.02.031.Search in Google Scholar
ScienceDirect (2024). Photocatalytic-search, https://www.sciencedirect.com/search?qs=photocatalytic (Accessed 25 April 2024).Search in Google Scholar
Sesana, E., Gagnon, A.S., Ciantelli, C., Cassar, J., and Hughes, J.J. (2021). Climate change impacts on cultural heritage: a literature review. Wiley Interdiscip. Rev. Clim. Change 12: e710, https://doi.org/10.1002/wcc.710.Search in Google Scholar
Sham, L.J. and Schlüter, M. (1983). Density-functional theory of the energy gap. Phys. Rev. Lett. 51: 1888, https://doi.org/10.1103/PhysRevLett.51.1888.Search in Google Scholar
Shao, Y., Gan, Z., Epifanovsky, E., Gilbert, A.T.B., Wormit, M., Kussmann, J., Lange, A.W., Behn, A., Deng, J., Feng, X., et al. (2015). Advances in molecular quantum chemistry contained in the Q-Chem 4 program package. Mol. Phys. 113: 184–215, https://doi.org/10.1080/00268976.2014.952696.Search in Google Scholar
Shao, Y., de Ruiter, J.M., de Groot, H.J.M., and Buda, F. (2019). Photocatalytic water splitting cycle in a dye-catalyst supramolecular complex: ab initio molecular dynamics simulations. J. Phys. Chem. C 123: 21403–21414, https://doi.org/10.1021/acs.jpcc.9b06401.Search in Google Scholar
Shen, R., Ren, D., Ding, Y., Guan, Y., Ng, Y.H., Zhang, P., and Li, X. (2020). Nanostructured CdS for efficient photocatalytic H2 evolution: a review. Sci. China Mater. 63: 2153–2188, https://doi.org/10.1007/s40843-020-1456-x.Search in Google Scholar
Shilpent (2024). High pressure mercury lamp, https://shilpent.com/other-products/67-high-pressure-mercury-lamp.html (Accessed 1 Aug 2025).Search in Google Scholar
Soler, J.M., Artacho, E., Gale, J.D., García, A., Junquera, J., Ordejón, P., and Sánchez-Portal, D. (2002). The SIESTA method for ab initio order-N materials simulation. J. Phys.: Condens. Matter 14: 2745, https://doi.org/10.1088/0953-8984/14/11/302.Search in Google Scholar
Spiegelman, F., Tarrat, N., Cuny, J., Dontot, L., Posenitskiy, E., Martí, C., Simon, A., and Rapacioli, M. (2020). Density-functional tight-binding: basic concepts and applications to molecules and clusters. Adv. Phys. X 5: 1710252, https://doi.org/10.1080/23746149.2019.1710252.Search in Google Scholar PubMed PubMed Central
Stark, W.G., Westermayr, J., Douglas-Gallardo, O.A., Gardner, J., Habershon, S., and Maurer, R.J. (2023). Machine learning interatomic potentials for reactive hydrogen dynamics at metal surfaces based on iterative refinement of reaction probabilities. J. Phys. Chem. C 127: 24168–24182, https://doi.org/10.1021/acs.jpcc.3c06648.Search in Google Scholar PubMed PubMed Central
Statista (2024). Global large-scale carbon capture projects, https://www.statista.com/statistics/726624/large-scale-carbon-capture-and-storage-projects-worldwide-by-status (Accessed 22 March 2024).Search in Google Scholar
Steane, A. (1998). Quantum computing. Rep. Prog. Phys. 61: 117, https://doi.org/10.1088/0034-4885/61/2/002.Search in Google Scholar
Stojkovic, M., Linscott, E., and Marzari, N. (2024). Predicting the suitability of photocatalysts for water splitting using Koopmans spectral functionals: the case of TiO2 polymorphs. arXiv preprint arXiv: 2412.17488.Search in Google Scholar
Subramanian, P., Meziane, D., Wojcieszak, R., Dumeignil, F., Boukherroub, R., and Szunerits, S. (2018). Plasmon-induced electrocatalysis with multi-component nanostructures. Mater. 12: 43, https://doi.org/10.3390/ma12010043.Search in Google Scholar PubMed PubMed Central
Takata, T., Jiang, J., Sakata, Y., Nakabayashi, M., Shibata, N., Nandal, V., Seki, K., Hisatomi, T., and Domen, K. (2020). Photocatalytic water splitting with a quantum efficiency of almost unity. Nature 581: 411–414, https://doi.org/10.1038/s41586-020-2278-9.Search in Google Scholar PubMed
Tang, D., Ketkaew, R., and Luber, S. (2024). Machine learning interatomic potentials for heterogeneous catalysis. Chem. Eur J. 30: e202401148, https://doi.org/10.1002/chem.202401148.Search in Google Scholar PubMed
Tarighati Sareshkeh, A., Seyed Dorraji, M.S., Karami, Z., Shahmoradi, S., Fekri, E., Daneshvar, H., Rasoulifard, M.H., and Karimov, D.N. (2023). Preparation of high-crystalline and non-metal modified g-C3N4 for improving ultrasound-accelerated white-LED-light-driven photocatalytic performances. Sci. Rep. 13: 15079, https://doi.org/10.1038/s41598-023-41473-y.Search in Google Scholar PubMed PubMed Central
Tian, L., Guan, X., Zong, S., Dai, A., and Qu, J. (2023). Cocatalysts for photocatalytic overall water splitting: a mini review. Catal. 13: 355, https://doi.org/10.3390/catal13020355.Search in Google Scholar
Tolod, K.R., Hernández, S., Castellino, M., Deorsola, F.A., Davarpanah, E., and Russo, N. (2020). Optimization of BiVO4 photoelectrodes made by electrodeposition for sun-driven water oxidation. Int. J. Hydrogen Energy 45: 605–618, https://doi.org/10.1016/j.ijhydene.2019.10.236.Search in Google Scholar
University College London (2024). Solar simulator for photoelectrochemical water splitting, https://www.ucl.ac.uk/solar-energy-advanced-materials/solar-driven-water-splitting (Accessed 22 March 2024).Search in Google Scholar
Ursua, A., Gandia, L.M., and Sanchis, P. (2011). Hydrogen production from water electrolysis: current status and future trends. Proc. IEEE 100: 410–426, https://doi.org/10.1109/JPROC.2011.2147110.Search in Google Scholar
U.S. Department of Energy (2023). DOE technical targets for hydrogen production: photoelectrochemical water splitting, https://www.energy.gov/eere/fuelcells/doe-technical-targets-hydrogen-production-photoelectrochemical-water-splitting (Accessed 1 Aug 2025).Search in Google Scholar
Vamvasakis, I., Andreou, E.K., and Armatas, G.S. (2023). Mesoporous dual-semiconductor ZnS/CdS nanocomposites as efficient visible light photocatalysts for hydrogen generation. Nanomater. 13: 2426, https://doi.org/10.3390/nano13172426.Search in Google Scholar PubMed PubMed Central
Vasilchenko, D., Zhurenok, A., Saraev, A., Gerasimov, E., Cherepanova, S., Kovtunova, L., Tkachev, S., and Kozlova, E. (2022). Platinum deposition onto g-C3N4 using a labile nitratocomplex for highly active hydrogen evolution photocatalysts. Int. J. Hydrogen Energy 47: 11326–11340, https://doi.org/10.1016/j.ijhydene.2021.09.253.Search in Google Scholar
Vedavyasa, K.V. and Kumar, A. (2024). Classification analysis of transition-metal chalcogenides and oxides using quantum machine learning. arXiv preprint arXiv: 2405.18989.10.1002/qute.202400081Search in Google Scholar
Villa, K., Galán-Mascarós, J.R., López, N., and Palomares, E. (2021). Photocatalytic water splitting: advantages and challenges. Sustain. Energy Fuels 5: 4560–4569, https://doi.org/10.1039/D1SE00808K.Search in Google Scholar
Walter, M.G., Warren, E.L., McKone, J.R., Boettcher, S.W., Mi, Q., Santori, E.A., and Lewis, N.S. (2010). Solar water splitting cells. Chem. Rev. 110: 6446–6473, https://doi.org/10.1021/cr1002326.Search in Google Scholar PubMed
Wang, F., Di Valentin, C., and Pacchioni, G. (2012). Rational band gap engineering of WO3 photocatalyst for visible-light water splitting. ChemCatChem 4: 476–478, https://doi.org/10.1002/cctc.201100374.Search in Google Scholar
Wang, Y., Ping, H., Tan, T., Wang, W., Ma, P., and Xie, H. (2019). Enhanced hydrogen evolution from water splitting based on ZnO nanosheet/CdS nanoparticle heterostructures. RSC Adv. 9: 28165–28170, https://doi.org/10.1039/C9RA04975D.Search in Google Scholar PubMed PubMed Central
Wang, Z., Lin, Z., Shen, S., Zhong, W., and Cao, S. (2021). Advances in designing heterojunction photocatalytic materials. Chin. J. Catal. 42: 710–730, https://doi.org/10.1016/S1872-2067(20)63698-1.Search in Google Scholar
Wang, H., Qi, H., Sun, X., Jia, S., Li, X., Miao, T.J., Xiong, L., Wang, S., Zhang, X., Liu, X., et al.. (2023a). High quantum efficiency of hydrogen production from methanol aqueous solution with PtCu–TiO2 photocatalysts. Nat. Mater. 22: 619–626, https://doi.org/10.1038/s41563-023-01519-y.Search in Google Scholar PubMed
Wang, K., Huang, D., Li, X., Feng, K., Shao, M., Yi, J., He, W., and Qiao, L. (2023b). Unconventional strategies to break through the efficiency of light-driven water splitting: a review. Electron 1: e4, https://doi.org/10.1002/elt2.4.Search in Google Scholar
Wang, Q., Zhu, Q., Cao, L., Fan, L., Gu, F., Zhang, Y., Zheng, C., Xiong, S., and Xu, L. (2024a). Two-dimensional SiH/g-C3N4 van der Waals type-II heterojunction photocatalyst: a new effective and promising photocatalytic material. Coatings 14: 263, https://doi.org/10.3390/coatings14030263.Search in Google Scholar
Wang, Y., Sorkun, M.C., Brocks, G., and Er, S. (2024b). ML-aided computational screening of 2D materials for photocatalytic water splitting. J. Phys. Chem. Lett. 15: 4983–4991, https://doi.org/10.1021/acs.jpclett.4c00425.Search in Google Scholar PubMed
Wang, C., Wang, B., Wang, C., Li, A., Chang, Z., and Wang, R. (2025). A machine learning model with minimized feature parameters for multi-type hydrogen evolution catalyst prediction. npj Comput. Mater. 11: 111, https://doi.org/10.1038/s41524-025-01607-4.Search in Google Scholar
Watkins, C.J.C.H. and Dayan, P. (1992). Q-learning. Mach. Learn. 8: 279–292, https://doi.org/10.1007/BF00992698.Search in Google Scholar
Wayo, D.D.K., Irawan, S., Satyanaga, A., and Kim, J. (2023). Data-driven fracture morphology prognosis from high-pressured modified proppants based on stochastic-Adam-RMSProp optimizers; tf.nnr study. Big Data Cogn. Comput. 7: 57, https://doi.org/10.3390/bdcc7020057.Search in Google Scholar
Wayo, D.D.K., Irawan, S., Satyanaga, A., Kim, J., Bin Mohamad Noor, M.Z., and Rasouli, V. (2024). Filter cake neural-objective data modeling and image optimization. Symmetry 16: 1072, https://doi.org/10.3390/sym16081072.Search in Google Scholar
Wexler, R.B., Martirez, J.M.P., and Rappe, A.M. (2018). Chemical pressure-driven enhancement of the hydrogen evolving activity of Ni2P from nonmetal surface doping interpreted via machine learning. J. Am. Chem. Soc. 140: 4678–4683, https://doi.org/10.1021/jacs.8b00947.Search in Google Scholar PubMed
Wheeler, T. and Von Braun, J. (2013). Climate change impacts on global food security. Sci. 341: 508–513, https://doi.org/10.1126/science.1239402.Search in Google Scholar PubMed
Winther, K., Hoffmann, M.J., Boes, J.R., Mamun, O., Bajdich, M., and Bligaard, T. (2019). Catalysis-Hub.org, an open electronic structure database for surface reactions. Sci. Data 6: 75, https://doi.org/10.1038/s41597-019-0081-y.Search in Google Scholar PubMed PubMed Central
Wu, X., Chen, G., Wang, J., Li, J., and Wang, G. (2023). Review on S-scheme heterojunctions for photocatalytic hydrogen evolution. Acta Phys.-Chim. Sin. 39: 221201, https://doi.org/10.3866/pku.whxb202212016, https://www.whxb.pku.edu.cn/EN/article/downloadArticleFile.do?attachType=PDF&id=37368 (Accessed 31 Jul 2025).Search in Google Scholar
Xie, Y., Mai, W., Ke, S., Zhang, C., Chen, Z., Wang, X., Zhu, S., Shen, Z., Zheng, W., Li, G., et al. (2024). Artificial intelligence-implemented prediction and cost-effective optimization of micropollutant photodegradation using g-C3N4/Bi2O3 heterojunction. Chem. Eng. J. 499: 156029, https://doi.org/10.1016/j.cej.2024.156029.Search in Google Scholar
Xu, P., Ji, X., Li, M., and Lu, W. (2023). Small data machine learning in materials science. npj Comput. Mater. 9: 42, https://doi.org/10.1038/s41524-023-01000-z.Search in Google Scholar
Xu, Y., Liu, X., Cao, X., Huang, C., Liu, E., Qian, S., Liu, X., Wu, Y., Dong, F., Qiu, C.W., et al. (2021). Artificial intelligence: a powerful paradigm for scientific research. Innovation 2, https://doi.org/10.1016/j.xinn.2021.100179, https://doi.org/100179.Search in Google Scholar PubMed PubMed Central
Yang, X., Zhao, R., Zhan, H., Zhao, H., Duan, Y., and Shen, Z. (2024). Modified titanium dioxide-based photocatalysts for water treatment: mini review. Environ. Funct. Mater. 3: 1–12, https://doi.org/10.1016/j.efmat.2024.07.002.Search in Google Scholar
Yoon, N., Jung, W., and Kim, H. (2024). Unveiling hidden insights in gas chromatography data analysis with generative adversarial networks. Chemosens. 12: 131, https://doi.org/10.3390/chemosensors12070131.Search in Google Scholar
Yu, D. and Cao, F. (2025). Construction and approximation rate for feedforward neural network operators with sigmoidal functions. J. Comput. Appl. Math. 453: 116150.10.1016/j.cam.2024.116150Search in Google Scholar
Yue, M., Lambert, H., Pahon, E., Roche, R., Jemei, S., and Hissel, D. (2021). Hydrogen energy systems: a critical review of technologies, applications, trends and challenges. Renew. Sustain. Energy Rev. 146: 111180.10.1016/j.rser.2021.111180Search in Google Scholar
Zhang, Y., Li, Y., Xin, X., Wang, Y., Guo, P., Wang, R., Wang, B., Huang, W., Sobrido, A.J., and Li, X. (2023). Internal quantum efficiency higher than 100% achieved by combining doping and quantum effects for photocatalytic overall water splitting. Nat. Energy 8: 504–514.10.1038/s41560-023-01242-7Search in Google Scholar
Zhang, C. and Lu, Y. (2021). Study on artificial intelligence: the state of the art and future prospects. J. Ind. Inf. Integr. 23: 100224.10.1016/j.jii.2021.100224Search in Google Scholar
Zhang, Q., Yue, M., Chen, P., Ren, Q., Kong, W., Jia, C., Lu, Q., Wu, J., Li, Y., Liu, W., et al. (2024). Accelerating photocatalytic hydrogen production by anchoring Pt single atoms on few-layer gC3N4 nanosheets with Pt–N coordination. J. Mater. Chem. C 12: 3437–3449.10.1039/D3TC04673GSearch in Google Scholar
Zhang, H., Zhang, B., Wang, X., Zou, L., You, J., and Lin, S. (2024). Effective charge separation in photoelectrochemical water splitting: a review from advanced evaluation methods to materials design. Sustain. Energy Fuel 8: 2357–2382.10.1039/D4SE00390JSearch in Google Scholar
Zhang, R., Zhao, J., Yang, Y., Lu, Z., and Shi, W. (2016). Understanding electronic and optical properties of La and Mn co-doped anatase TiO2. Comput. Condens. Matter 6: 5–17.10.1016/j.cocom.2016.03.001Search in Google Scholar
Zhao, F., Zhang, D., Ni, Y., Wang, H., and Peng, S. (2025). Tunable band gaps and conduction band edges of CdS/ZnS heterostructures – a first-principles-based prediction. Phys. Chem. Chem. Phys. 27: 1852–1860.10.1039/D4CP03827DSearch in Google Scholar PubMed
Zhou, Q., Lian, Y., Wu, J., Zhu, M., Wang, H., and Cao, J. (2024). An optimized Q-Learning algorithm for mobile robot local path planning. Knowl.-Based Sys. 286: 111400.10.1016/j.knosys.2024.111400Search in Google Scholar
Zhou, J., Liu, Q., Feng, W., Sun, Y., and Li, F. (2015). Upconversion luminescent materials: Advances and applications. Chem. Rev. 115: 395–465.10.1021/cr400478fSearch in Google Scholar PubMed
Zhou, P., Navid, I.A., Ma, Y., Xiao, Y., Wang, P., Ye, Z., Zhou, B., Sun, K., and Mi, Z. (2023). Solar-to-hydrogen efficiency of more than 9% in photocatalytic water splitting. Nature 613: 66–70.10.1038/s41586-022-05399-1Search in Google Scholar PubMed
Zhu, X., Lin, Y., Sun, Y., Beard, M.C., and Yan, Y. (2019). Lead-halide perovskites for photocatalytic $\alpha$-alkylation of aldehydes. J. Am. Chem. Soc. 141: 733–738.10.1021/jacs.8b08720Search in Google Scholar PubMed
© 2025 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Frontmatter
- Reviews
- Advancing hydrogen infrastructure: a review of storage and transportation solutions for a sustainable future
- DFT and hybrid classical–quantum machine learning integration for photocatalyst discovery and hydrogen production
- Advancements in synthesis methods and their effects on the physico-chemical properties and yield efficiency of ZSM-5/SAPO-34 composites: a comprehensive review
Articles in the same Issue
- Frontmatter
- Reviews
- Advancing hydrogen infrastructure: a review of storage and transportation solutions for a sustainable future
- DFT and hybrid classical–quantum machine learning integration for photocatalyst discovery and hydrogen production
- Advancements in synthesis methods and their effects on the physico-chemical properties and yield efficiency of ZSM-5/SAPO-34 composites: a comprehensive review