Next Article in Journal
SCC Susceptibility of Polystyrene/TiO2 Nanocomposite-Coated Thin-Sheet Aluminum Alloy 2024—T3 in 3.5% NaCl
Previous Article in Journal
Shock Thermal Resistance of Parachute Fabrics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Water Electrolysis Technologies and Their Modeling Approaches: A Comprehensive Review

by
Ajitanshu Vedrtnam
1,2,*,
Kishor Kalauni
2 and
Rahul Pahwa
2
1
Mineral and Energy Economy Research Institute, Polish Academy of Sciences, Wybickiego 7A, 31-343 Kraków, Poland
2
Institute of Engineering and Technology, Invertis University, Bareilly 243123, India
*
Author to whom correspondence should be addressed.
Submission received: 5 March 2025 / Revised: 15 April 2025 / Accepted: 17 April 2025 / Published: 21 April 2025

Abstract

:
Hydrogen (H2) is a key energy vector in the global transition toward clean and sustainable energy systems. Among the various production methods, water electrolysis presents a promising pathway for zero-emission hydrogen generation when powered by renewables. This review provides a comprehensive evaluation of water electrolysis technologies, including alkaline (AWE), proton exchange membrane (PEMWE), solid oxide (SOEC), anion exchange membrane (AEMWE), and microbial electrolysis cells (MEC). It critically examines their material systems, catalytic strategies, operational characteristics, and recent performance advances. In addition to reviewing experimental progress, the study presents a finite element modeling (FEM) case study that evaluates thermal and mechanical responses in PEM and AWE configurations—illustrating how FEM supports design optimization and performance prediction. To broaden methodological insight, other simulation frameworks such as computational fluid dynamics (CFD), response surface methodology (RSM), and system-level modeling (e.g., Aspen Plus®) are also discussed based on their use in recent literature. These are reviewed to guide future integration of multi-scale and multi-physics approaches in electrolyzer research. By bridging practical design, numerical simulation, and material science perspectives, this work provides a resource for researchers and engineers advancing next-generation hydrogen production systems.

Graphical Abstract

1. Introduction

The global demand for sustainable energy solutions has led to an increased focus on hydrogen as a promising energy carrier due to its high energy density and potential for zero-emission applications [1,2]. While hydrogen production is currently dominated by fossil fuel-based methods such as steam methane reforming and coal gasification, these processes contribute significantly to carbon emissions [3]. In contrast, water electrolysis presents a cleaner alternative, enabling hydrogen generation without direct greenhouse gas emissions when powered by renewable energy sources. However, the widespread adoption of electrolysis technologies is hindered by challenges related to energy efficiency, system design, and material durability, necessitating advancements in optimization strategies and predictive modeling. Efficient hydrogen production through electrolysis is only one part of the solution; equally important is the ability to store and utilize this energy effectively, making advancements in energy storage technologies crucial for the widespread adoption of hydrogen as a sustainable fuel.
Energy storage involves converting complex energy forms into more efficient and economical storage solutions, playing a crucial role in economic development. Recent advancements, particularly in nanomaterials for ultracapacitors, have significantly improved battery lifespan and capacity. Various energy storage technologies have been developed, as outlined in Table 1 [4].
Among all, the storage and production of H2 are incredibly significant nowadays because of its high energy capacity and transportability [5]. Hydrogen is a widely available energy carrier that can be produced from both renewable and non-renewable sources. It offers high energy density, sustainability, and storage for renewable energy derived electricity, with water as its only byproduct and no carbon emissions [6]. Hydrogen can be produced through various methods at different stages of development (Figure 1). Currently, 78% comes from fossil fuel reforming, 14% from coal gasification, and only 4% from alternative sources, primarily electrolysis [4].
Water electrolysis, first reported in 1789, is an electrochemical process that uses electrical energy to split water into hydrogen and oxygen [7]. It consists of a water-filled cell with externally powered electrodes, where hydrogen forms at the negative electrode and oxygen at the positive. Efficiency can be optimized by adjusting electrolysis type, pressure, temperature, current density, electrolytes, membranes, electrodes, and catalysts [8]. The variation in factors and their effect on electrolysis are summarized in Table 2 [9,10,11].
Alkaline, polymer electrolyte membrane, solid oxide, and microbial electrolysis cells have been widely studied, each with unique advantages and challenges. Research has improved electrode materials, catalysts, electrolytes, and operating conditions, yet key issues persist, including energy losses from overpotential, internal resistance, and mass transport limitations. Electrode degradation, membrane instability, and gas bubble formation further reduce efficiency and increase energy consumption, necessitating systematic optimization [10,11]. Computational modeling and finite element analysis are crucial for optimizing electrolysis by providing insights into electrochemical kinetics, ion transport, thermal stability, and mechanical integrity. Simulations such as computational fluid dynamics and multi-physics modeling enable better system design, yet a gap remains in integrating numerical models with experimental validation, limiting predictive accuracy and scalability.
This review addresses these gaps by providing a comprehensive assessment of advancements in water electrolysis technologies while emphasizing the role of numerical modeling in optimizing performance. It critically evaluates different electrolysis methods in terms of operational efficiency, material selection, and system constraints while highlighting computational approaches that enhance process predictability and efficiency. By integrating insights from both experimental studies and numerical modeling, this review aims to establish a framework for improving water electrolysis technologies, bridging the gap between theoretical predictions and practical implementation. The findings presented herein contribute to the ongoing efforts to enhance the scalability and economic viability of hydrogen production through electrolysis, paving the way for future research in energy-efficient hydrogen generation.

2. Methodology

This review explores water electrolysis technologies, emphasizing the optimization and efficiency improvements in hydrogen production, with attention to electrode materials, catalyst advancements, energy consumption, and renewable energy integration. Initially, an extensive literature survey was conducted, revealing a lack of studies that integrate these aspects of electrolysis into a single review. Data were collected from a wide range of sources, including research articles, original works, books, reviews, proceedings, and dissertations, ensuring a thorough and well-rounded perspective. A systematic review was conducted following the preferred reporting items for systematic reviews and meta-analyses (PRISMA) methodology (Figure 2), which involves a detailed and transparent approach to article selection, search strategy, data extraction, and data analysis. The PRISMA process ensures a structured and transparent approach to systematically identifying and analyzing relevant studies, enabling the consolidation of key findings across diverse electrolysis technologies. By applying this methodology, the review effectively highlights critical gaps in the optimization of electrolysis processes, efficiency enhancements, and cost reduction strategies, paving the way for addressing unexplored synergies in sustainable hydrogen generation.
A thorough search was conducted across leading online databases, including SCI, SCIE, Elsevier Scopus, and Web of Science, encompassing a wide range of scientific disciplines. Only peer-reviewed studies from high-impact journals indexed in SCI, SCIE, Scopus, and Web of Science were selected to ensure data reliability. The selection of articles was carried out in three stages. The first stage, initial identification, involved inputting search terms aligned with the research objectives into relevant databases to identify potentially relevant studies. In the second stage screening, studies were evaluated based on (i) relevance to the research objectives, including system optimization, material development, performance analysis, or modeling of water electrolysis; (ii) scientific quality, assessed through methodological clarity, completeness of reporting, peer-reviewed status, and validation or reproducibility of data; (iii) study design, encompassing both experimental and computational work as well as systematic reviews; and (iv) language, with only articles published in English considered. The final stage, full-text review, involved a comprehensive examination of the remaining studies to ensure they fully met all eligibility requirements.
A wide range of keywords, including “water electrolysis”, “hydrogen production”, “efficiency”, “optimization”, “catalyst”, “electrode materials”, “nanomaterials”, “renewable energy integration”, “cost reduction”, “proton exchange membrane (PEM) electrolysis”, “alkaline water electrolysis (AWE)”, “solid oxide electrolysis cell (SOEC)”, “anion exchange membrane (AEM) electrolysis”, “high-temperature electrolysis”, and “low-temperature electrolysis”, were employed to construct this review paper. Keyword searches were performed using Boolean operators, combining terms such as ‘water electrolysis’ AND ‘hydrogen production’ OR ‘efficiency improvement’ to refine results. The search was repeated across different databases to ensure comprehensive coverage. The selection of publications prioritized factors such as the reliability and validity of data, recency of research, relevance, impact factors, citation count, and utility of the articles. Data collection was conducted manually through content analysis, extracting critical information such as article type, journal name, year of publication, topic, title, research methodology, relationships between variables, indicators, and research findings. The articles retained after the final screening provided a comprehensive foundation for evaluating the synergies between electrolysis technologies, energy efficiency improvements, and cost-effective hydrogen production.

3. Classification and Analysis of Water Electrolysis Technologies

Figure 3, Figure 4 and Figure 5 present a classification of the literature on various aspects of water electrolysis technologies. Electrolysis is categorized into alkaline water electrolysis (AWE), solid oxide electrolysis cells (SOEC), polymer electrolyte membrane (PEM) electrolysis, and microbial electrolysis cells (MEC), with AWE being the most extensively studied. Further classification is based on electrode materials, where nickel-based and oxide electrodes are predominantly used for both anode and cathode applications. Research also classifies electrolysis studies by current density variations, with most investigations utilizing high current densities to enhance hydrogen production. Additionally, studies differentiate electrolysis based on electrolyte type, with SSZ, barium cerate carbonate, and potassium hydroxide (KOH) being the most commonly used, while yttria-stabilized zirconia (YSZ) and sodium hydroxide (NaOH) appear less frequently. Electrolysis has been performed using various water sources, including pure water, steam, wastewater, and seawater, with researchers also evaluating the impact of temperature on system performance.
Figure 3a illustrates that 30% of electrolysis studies focus on AWE. Figure 3b,c highlight the predominance of nickel-based and oxide materials as electrode choices, while Figure 3d confirms the frequent use of SSZ, barium cerate carbonate, and KOH as electrolytes. These classifications provide valuable insights into trends in electrolysis research, guiding advancements in material selection and process optimization.
Figure 4a classifies electrolysis studies based on membrane type, with titanium porous transport layers, Nafion 115 membranes, and co-crystallized catalyst-coated membranes being the most commonly used. Figure 4b highlights the frequent use of platinum-based catalysts, indicating their widespread preference in electrolysis research.
Similarly, Figure 5a classifies electrolysis studies based on the state of H2O at the inlet, with steam (36%) and wastewater (36%) being the most frequently used, while pure water is also widely studied. Figure 5b provides a breakdown of electrolysis research based on the operating temperature ranges used in various studies. The chart indicates that 75% of the research focuses on high temperature electrolysis (>500 °C), underscoring the crucial role of elevated temperatures in improving process efficiency. High temperatures facilitate faster reaction kinetics, lower the energy barrier for water splitting, and improve overall system performance, making it the most extensively studied condition. In contrast, medium-temperature electrolysis (200–500 °C) accounts for only 12% of the studies, and low temperature electrolysis (<200 °C) represents 13%. The smaller percentages suggest that while lower-temperature systems are explored, they present challenges such as higher overpotentials and lower reaction rates, which could limit their efficiency and applicability in large-scale operations. Overall, the data reflects the strong research interest in high temperature electrolysis, likely driven by its potential for enhanced hydrogen production rates, improved system integration with industrial heat sources, and better economic viability over the long-term.

3.1. Alkaline Water Electrolysis (AWE)

In AWE (Figure 6a), electrodes are immersed in an alkaline solution and separated by a membrane, which acts as a diaphragm to prevent gas crossover while facilitating the transport of hydroxide ions (OH) between electrodes, ensuring efficient electrochemical reactions [12]. The reactions, charge carriers, and overall cell reactions occurring in AWE are detailed in Table 3. As a commercially viable and cost-effective technology, AWE benefits from the use of non-noble electrocatalysts, making it an attractive option for large-scale hydrogen production. However, it also faces certain limitations, including low current densities, carbonate formation, reduced gas purity, and low operating pressures, which impact overall performance. The energy efficiency of AWE typically ranges between 70 and 80%, with further improvements dependent on optimizing operating conditions and material advancements [13]. The efficiency of AWE can be optimized by adjusting current density, electrode materials, electrolyte composition, and membrane properties, all of which significantly influence overall performance and hydrogen production rates.
Researchers are actively exploring innovative approaches to enhance the efficiency and sustainability of hydrogen production through electrolysis. Bhattacharya et al. advocated for the integration of solar energy with AWE to support a sustainable hydrogen economy [14]. Bidin et al. demonstrated that using collimated light increased hydrogen production by 53% compared to conventional light and dark field conditions [15]. Giraldi et al. utilized nuclear power for electrolysis, reporting low greenhouse gas emissions of 416 g CO2eq per kg H2 [16]. Bhandari et al. emphasized the preference for hydropower or wind-generated electricity over fossil fuels for electrolysis-based hydrogen production [17]. Kwak et al. developed a W-typed dye-sensitized serial solar module to improve H2 generation efficiency [18]. Kelly et al. highlighted the role of electrolytic hydrogen in fuel-cell electric vehicles, enhancing efficiency while reducing environmental impact [19]. Wang et al. suggested that applying an external field and modifying electrolyte composition could mitigate reaction overpotential and reduce ohmic voltage drop, further improving electrolysis performance [20].
AWE has emerged as one of the most mature and scalable technologies for green hydrogen production due to its reliance on non-noble metals and compatibility with renewable energy sources. Xie and Shao [21] emphasize the commercial viability of AWE and its relatively low capital cost compared to PEM and SOEC systems, despite inherent challenges such as sluggish electrode kinetics and system inefficiencies. Tüysüz [22] further elaborates on these kinetic challenges, particularly highlighting the need for improved oxygen evolution reaction (OER) catalysts, given its thermodynamic and kinetic complexity. In a modeling centric approach, Hu et al. [23] argue that comprehensive thermodynamic, electrochemical, and thermal models are essential to optimize operational efficiency and improve system-level design for AWE. Ehlers et al. [24] underscore the importance of bridging the gap between lab-scale innovation and industrial scale deployment by focusing on durability, cost, and system integration. Thissen et al. [25] reinforce this by advocating for lab-scale protocols that replicate industrial conditions—such as high temperatures and concentrated KOH electrolytes—to accurately assess catalyst performance. Dou et al. [26] use scanning acoustic microscopy to visualize two-phase flow dynamics in porous electrodes, providing actionable insights into bubble management and electrolyte replenishment—factors critical for high current density performance. Brauns and Turek [27] explore the integration of AWE with intermittent renewable sources such as wind and solar, noting the need for dynamic operation and optimized part-load performance to prevent safety hazards and maximize efficiency. Finally, Sebbahi et al. [28] provide a holistic overview of electrolysis technologies, reaffirming AWE as the most cost-effective and industrially ready solution, despite the need for continued research in catalyst durability and system optimization.
The efficiency of AWE varies with different cathode materials, with Ni-based materials being the most widely used. Kabulska et al. investigated Ni-Fe-C alloys due to their high electroactivity for the hydrogen evolution reaction (HER) in AWE. By applying plasma treatment using CH4 and H2 at 470 °C, they successfully introduced carbon into Ni and Ni-Fe alloys, leading to enhanced catalytic activity. This improvement was attributed to carburization, which significantly boosted HER performance in 25 wt.% KOH at 80 °C [29]. This carburization not only improved catalytic activity by increasing surface reactivity but also improved electrode durability under alkaline conditions. Sivanantham et al. integrated Ni3Se2 with Ni-foam as a cathode structure, enhancing electron conductivity and active surface area. The foam served as a high surface area support, promoting gas release and improving hydrogen production continuity during long-term operation [30]. Solmaz et al. engineered Cu/Ni/NiZn-PtRu multilayer cathodes to leverage synergistic catalytic effects across metal layers. The PtRu addition enhanced HER kinetics, while the NiZn alloy improved corrosion resistance, making this structure promising for prolonged AWE cycling [31]. Furthermore, Buch et al. demonstrated that incorporating Au nanoparticles into Ni electrodes significantly improved HER performance by increasing the electrochemical surface area and modifying electronic properties at the metal interface, thereby reducing the activation overpotential [32]. Rauscher et al. investigated HER of nanocrystalline NiMoB alloys in 1 M KOH to compare the equivalent crystalline and polycrystalline materials at 298 K [33]. Kim et al. demonstrated a low-cost, asymmetric porous nickel electrode for alkaline water electrolysis by sintering nickel powder into nickel foam, creating a dual surface architecture. The fine-pore surface acted as the electro-catalytic interface, while the open pore side facilitated gas and electrolyte transport. This design achieved high current densities (~0.5 A/cm2 at 1.8 V, 80 °C), enhanced mass transfer, and long-term operational stability over 400 h. The integration of lightweight gaskets and polymer separators into a compact stack further highlighted the system’s scalability and energy efficiency [34]. Kabulska et al. applied plasma carburization to pure nickel cathodes using CH4 + H2 gas mixture at 470 °C, resulting in carbon-enriched surfaces. The treated electrodes exhibited enhanced catalytic activity for HER in 25 wt.% KOH at 80 °C, particularly in early stages of operation. The improvement was attributed to surface-bound carbon increasing electrocatalytic activity rather than surface area effects. Performance gains diminished after prolonged operation due to iron deposition from the electrolyte [35]. Cardoso et al. prepared nickel-samarium and nickel-dysprosium alloys of rare earth metal with 5% and 10% by arc and induction furnace melting, respectively. The results indicated that the electrocatalytic performance of Ni-Re alloys with 5% of rare earth metal for HER is better than 10% of rare earth metal [36]. Park et al. have reported the use of quaternary ammonium-tethered aromatic polymers for anion exchange membrane-based AWE [37]. Tang et al. reported a fabrication procedure of the Ni-Mo electrodes for HER. The Ni-Mo electrodes exhibited a high catalytic activity in KOH solution at 70 °C towards HER [38].
A number of other materials other than nickel are also reported as cathode material during AWE. Zuo et al. [39] address these concerns by introducing Ru-perturbed Cu nanoplatelet cathodes that achieve current densities up to 3.6 A/cm2 at 2 V, rivaling PEM electrolyzers while minimizing precious metal use. Yuksel et al. [40] fabricated and characterized 3D silver nanodomes by soft and nanosphere lithography. A higher rate of hydrogen production was observed for 3D AgNDs in 6 M KOH solution as opposed to Ag bulk [41]. Furthermore, Allebrod et al. used Co- and Mo-activated electrodes in AWE at 150–250 °C and 40 bars. It was observed that with the addition of cobalt oxide and molybdenum oxide electrocatalysts, the performance of the electrolysis cells was improved. Liu et al. [42] contribute by developing a Co3Mo3N/Co4N/Co heterostructure catalyst that promotes rapid electron transport and robust HER and OER kinetics under alkaline conditions. While the structure also exhibited OER potential, their study primarily highlights its cathodic role in AWE. Moreover, Muller et al. on the basis of their investigation on amorphous Fe-based alloys (Fe82B18, Fe80Si10B10, and Fe80Co20Si10B10) for AWE, documented that the addition of metalloids, especially cobalt, has been seen to enhance the HER activity of the materials [43]. Shen et al. have performed electrolysis using bauxite electrodes and using NaOH solution as the supporting electrolyte. The results indicated the desulfurization rate, the pH value, the electrode corrosion, and the cell voltage of the electrolyte [44]. Studies of Yao et al. revealed that fabricating Ni_Fe with Co and Ti as substrate (100 mA cm−2 for 100 h) showed extensive stability with minimal distortion of electrodes [45]. Even at low current densities, Ag nanodomes (AgNDs) exhibit a higher potential difference compared to bulk Ag, demonstrating their superior electrochemical performance.
Luo et al. [46] demonstrated significant improvements in the cathode of the alkaline water electrolyzer. By employing a nickel-plated nickel mesh with surface activation, they enhanced the electrochemically active surface area, effectively reducing overpotential and improving hydrogen production efficiency. The mesh structure not only increased surface area but also improved bubble release, contributing to better system performance. Egert et al. [47] expanded upon this by showing how low-tortuosity porous nickel electrodes dramatically improve efficiency at industrially relevant current densities (up to 2 A/cm2 at ~2 V). Performance analysis revealed that the cathode modification (Mo-incorporated porous Ni) had a greater impact on improving overall cell performance, particularly at high current densities. Thus, although both electrodes were enhanced, the cathode played a more dominant role in the performance improvement. These materials reduce the capillary pressure and bubble point, thereby enhancing two-phase mass transport and lowering ohmic resistance. Notably, their study confirmed these results both in lab-scale cells and a 2 kW AWE stack, indicating that porosity and microstructure play vital roles in maintaining high performance under load. Xia et al. [48] also noted that cathode design is critical under low-load conditions, especially when AWE systems are driven by intermittent renewable energy sources. They highlighted that the physical structure, not just the chemical composition, of the cathode affects both efficiency and consistency. For instance, inconsistent operation across multiple cells was linked to uneven electrode characteristics. This insight pushes beyond material composition to suggest structural uniformity is equally important. In the broader context of electrolyzer technologies, Nasser et al. [49] emphasized that nickel and nickel-alloy-based cathodes are common features in AWEs due to their robustness and cost-efficiency. Their review highlighted that while AWE efficiency ranges from 62–82% HHV, this can be significantly optimized by tailoring cathode materials and configurations. El-Shafie [5] provided a comparative overview of water electrolysis technologies, reiterating that AWE cathodes (typically nickel-based) offer a favorable balance between cost, efficiency, and operational lifetime. The efficiency losses in AWEs are often attributed to both ohmic and activation overpotentials, where the cathode’s electrocatalytic performance is central to minimizing these losses.
Zeng et al. enhanced the cathode performance of Ni-based electrodes through mechanical polishing with sandpaper and electrochemical deposition of Co and Ni. The polished electrodes exhibited superior apparent activity, achieving an overpotential of 422 mV at a current density of 750 A m−2, demonstrating significant improvement in HER. [50]. This study focuses on improving the cathode performance by surface modifications of Ni-based electrodes, enhancing the hydrogen evolution reaction (HER) through mechanical polishing and electrochemical deposition. Ahn et al. prepared the alloy catalyst of NiCu for AWE by changing the alloy composition with the electrodeposition method and investigated the catalytic activities in a 6.0 M KOH electrolyte at 298 K to HER with cyclic voltammetry [51]. Hong et al. prepared a NiW alloy catalyst on a copper foil substrate via a electrodeposition method for the HER in AWE. As observed, the morphologies of the NiW alloys exhibited severe variations as compared to Ni, at high W contents in the catalyst [52]. Egelund et al. have used the LaNiO3 electrode for OER in AWE. As a result, it was observed that the coating of LaNiO3 and Ni composite coatings decreased OER overpotential by 70 mV [53].
To consolidate the diverse approaches discussed for cathode optimization in AWE, Figure 7 illustrates a synthesized classification of material types, surface modification techniques, and their corresponding effects on electrochemical performance. Our synthesis (see Figure 7) visualizes these innovations alongside trade-offs and technical readiness levels to guide future material selection and device engineering. The framework emphasizes the interplay between structural enhancements (e.g., alloying, nano-structuring), performance outcomes (e.g., reduced overpotentials, enhanced current density), and operational considerations such as scalability, thermal compatibility, and material cost. This visual summary is intended to guide material selection by highlighting both innovation pathways and trade-offs relevant to practical implementation.
Apart from cathode development, significant work has also focused on improving anode catalysts for the OER. Jovic et al. investigated the OER in 1 M NaOH at 25 °C using pure Ni and Ni-Ebonex/Ir composite coatings. Initially, the Ni coating exhibited superior intrinsic catalytic activity for OER compared to the composite coatings. However, a significant loss of activity was observed after 24 h at 50 mA cm−2, indicating stability limitations in prolonged operation [54]. He et al. modified Ni anode by TiO2 nanotubes and found out that the hydrogen production rate was increased and the applied direct voltage reduced compared to normal water electrolysis [55].
Kuleshov et al. enhanced alkaline electrolysis performance by modifying porous nickel electrodes at both the anode and cathode. NiCo2O4 spinels were deposited onto the anode to reduce the OER overpotential, while NiPx or Pt nanoparticles were incorporated on the cathode to enhance HER kinetics. These dual modifications led to significant reductions in both overpotentials, ultimately lowering the total cell voltage and improving overall energy efficiency [56].
For large-scale hydrogen production via water electrolysis, the development of efficient, stable, and cost-effective OER catalysts is crucial. Li et al. synthesized an Fe-Co-Ni oxygen evolution catalyst via electrodeposition on nickel foam (NF), which exhibited high stability and enhanced catalytic activity, making it a promising candidate for improving the efficiency of water electrolysis [57]. Li et al. studied Pt-Mn2O4 catalysts with carbon black as a conductive agent and showed a remarkable activity for OER due to a synergistic effect between Mn2O4 and Pt. The Pt-Mn2O4/C catalyst with a weight ratio of Pt to Mn2O4 of 3:1 gives the best performance [58]. Gao et al. synthesized 2D FeSe2 nano-platelets via a hydrothermal reduction route, which exhibits extraordinarily high catalytic activities and stability for OER [59].
The effectiveness of AWE is also affected by the variation of current density (Figure 8). Dobo et al. emphasized the fact that there exists a relation between the efficacy of AWE and the variations of electric current. It was stated that at a constant current density, steady DC current causes much less efficiency loss as compared to the non-steady current conditions [60]. Increasing current density enhances hydrogen generation efficiency, while a higher ripple factor negatively impacts performance. Electrolyte composition also plays a crucial role in AWE efficiency. Berg et al. investigated potassium dihydrogen phosphate (KH2PO4) as an electrolyte and demonstrated that at 300 °C, it dissociates into H2O gas in equilibrium with KH2PO4, K2H2P2O7, and KPO3, highlighting its potential for improving electrolysis performance at elevated temperatures. As described in the study, at higher temperatures (around 275–325 °C), the Gibbs free energy (ΔG) of the water splitting reaction decreased by approximately 6%, making the electrolysis process thermodynamically more favorable and potentially more energy-efficient. Moreover, operating at elevated temperatures allowed the use of non-precious metal catalysts, which were less costly compared to the precious metals typically required at lower temperatures. In the study molten KH2PO4 acted as a stable electrolyte that retained significant water content through hydrogen bonding, enabling efficient high temperature electrolysis while containing water under high vapor pressure conditions [61]. Furthermore, Tufa et al. [62] fed a lab-scale reverse electrodialysis unit with dissimilar NaCl concentrations, imitating seawater and brine, to drive an alkaline polymer electrolyte water electrolysis cell. As shown in Figure 9b, the experimental setup includes gas separators at both electrodes and a salinity-gradient-driven RED stack with 27 cell pairs. Their system achieved 0.86 W power output, with the performance curve in Figure 9c showing a maximum gross power density under a temperature of 65 °C. This setup demonstrates the feasibility of integrating salinity-gradient energy with AWE to improve energy input efficiency. Nikiforov has also performed electrolysis using KH2PO4 at 300 °C. Hence, for temperatures as high as 300 °C, molten KH2PO4 as an electrolyte has shown promising results, depicting conductivity of ~0.30 S cm−1 at 300 °C [63]. On the contrary, Fiegenbaum et al. advocated the use of tetra-alkyl-ammonium-sulfonic acid as the electrolyte, which worked well even at room temperatures, producing current densities with high efficiencies [64]. Menia et al. explored methanol-assisted electrolysis as an alternative to conventional water electrolysis, using a 4 M aqueous methanol solution. This configuration replaces the OER at the anode with the methanol oxidation reaction (MOR), which proceeds at a lower overpotential. As a result, the system achieves hydrogen production at significantly lower voltages (0.65–0.85 V at 0.5–1 A/cm2), improving energy efficiency [65]. However, this process should be considered a distinct electrochemical pathway rather than a direct optimization of water electrolysis, since MOR alters the anodic chemistry and may produce CO2 depending on catalyst selectivity. Ge et al. used coal-water slurry and investigated the mechanism of electrolysis by anode reaction kinetics, whereby they demonstrated an inverse relation between activation energies of the electrode reaction and the activity of carbon materials [66]. Chang et al. stated that the Ni12P5/Ni3(PO4)2 hollow sphere exhibited high stability and activity towards OER and HER in alkaline media [67]. Li et al. [68] emphasized that modern AWEs are traditionally limited by low current density operation, often below 0.5 A/cm2, primarily due to bubble-induced overpotentials. Their study introduced a superaerophobic electrode design that significantly mitigated these effects, achieving an impressive current density of 3.5 A/cm2 at 2.25 V and 85 °C in a zero-gap configuration. The key to this performance enhancement was efficient gas bubble removal, which otherwise blocks active sites and increases ohmic losses at higher current densities. Jang et al. [69] conducted a comprehensive simulation study and demonstrated that increasing current density can enhance hydrogen production, but only within optimal temperature and pressure conditions. They reported that at 1 A/cm2, the AWE system achieved its highest efficiency of 78.52% at 120 °C and 10 bar. However, they also cautioned that excessive current density leads to thermal management challenges and reduced system stability due to increased hydrogen crossover risks. Ding et al. [70] provided a detailed exergy analysis and found that the ohmic overpotential varies more significantly with current density than activation overpotential. Their thermodynamic-electrochemical model showed that membrane resistance, which is sensitive to bubble formation and diaphragm properties, dominated the total ohmic losses under increasing current densities. This suggests that minimizing diaphragm resistance and optimizing bubble removal are crucial for maintaining efficiency at high loads.
The plot in Figure 9a shows the relationship between the partial pressure of H2O and temperature for KH2PO4. This was determined by Raman spectroscopy, using relative ratios of Raman scattering cross-sections of CH4 and H2O [41]. As the temperature increases, the partial pressure of water also rises, showing an exponential trend, especially above 200 °C. The melting point of KH2PO4 is marked at 270 °C, with an uncertainty of ±2 °C. This suggests that as the salt approaches and surpasses its melting point, the release of water becomes more significant, possibly due to phase change behavior and enhanced vaporization. Figure 9d shows a plot of the logarithm of conductivity versus inverse absolute temperature (1000/T) for different electrolyte molarities. This is a typical Arrhenius-type plot, where conductivity decreases with increasing 1000/T, indicating enhanced ion mobility at higher temperatures. Different symbols represent various electrolyte concentrations (e.g., 5 M Na3PO4, 3 M Na3PO4 + 2 M KH2PO4, etc.). Higher concentrations generally show higher conductivity, emphasizing the role of electrolyte composition in improving ionic transport. Figure 9e presents the polarization curves showing current density versus potential at varying electrolyte molarities (e.g., 0.1 M, 0.2 M, 0.3 M, etc.). As the applied potential becomes more negative, the current density increases—characteristic of cathodic reactions such as HER. Moreover, higher molarity electrolytes yield greater current densities at a given potential, reflecting reduced ohmic losses and improved ionic transport. This demonstrates the beneficial role of electrolyte concentration in enhancing the electrochemical performance of the system. Considering the fact that the membrane has a crucial role to play when it comes to efficiency in AWE, Aili et al. equilibrated the membrane in 22 wt.% aqueous KOH. This modified version of the porous poly(perfluorosulfonic acid) membrane was seen to deliver significantly higher ion conductivity (0.2 S/cm) even at room temperature, as opposed to the unmodified membrane (0.01 S/cm) [71].
For AWE, Burnat et al. made use of composite membranes integrated with polysulfone and mineral fillers. The proposed polysulfone-barite membranes are more advantageous than Zirfon 500 utp due to their lower ionic resistivity, higher hydrogen purity (99.83%), and significantly lower material cost. Additionally, barite offers excellent chemical stability in alkaline conditions, making these membranes both more efficient and cost-effective for industrial alkaline electrolysis [72]. Likewise, Wu et al. [73] review the evolution of porous membranes in AWE for green hydrogen production, emphasizing the transition from asbestos to safer non-asbestos alternatives. They highlight the role of materials such as PPS and Zirfon® in enhancing membrane efficiency, durability, and hydrophilicity. Diaz et al. have utilized alkali-doped poly (2,5 benzimidazole) membrane for poly (2,5-benzimidazole) membrane for AWE. It was reported that the linear and cross-linked membranes were stable up to 3.0 mol/dm3 and 4.2 mol/dm3 KOH doping, respectively [74]. In summary, AWE cathode optimization efforts have evolved from simple Ni-based systems to increasingly complex multi-metal and nanostructured materials. While many studies report improvements in HER kinetics or current density, the underlying mechanisms often involve enhanced bubble detachment, increased surface roughness, or improved electrical conductivity. However, cost and long-term stability remain critical bottlenecks for industrial scale-up.
The efficiency of AWE is also influenced by operating pressure (Figure 10). Suermann et al. studied the impact of pressures up to 100 bars on cell voltage behavior and observed that lowering the pressure reduced mass transport losses, thereby improving overall system efficiency [75]. Koi et al. analyzed the environmental impact of high pressure AWE systems, focusing on innovative membrane technologies. Their findings revealed that while improved membranes enhanced performance, the primary environmental impact stemmed from the electricity required for operation, emphasizing the need for renewable energy integration to minimize the carbon footprint [76]. Todd et al. have also performed electrolysis by increasing pressure (up to 100 MPa) at operating temperatures (up to 1000 K) and demonstrated lower energy requirements, albeit with electrical [77].
Furthermore, many researchers have reported the review and numerical modeling of AWE. To find out the impact of varied material parameters on the ohmic overpotential of AWE, Zouhri et al. developed a numerical and theoretical model that analyzes the contribution of ohmic polarization to overall energy losses in alkaline water electrolysis. This model evaluates the effects of membrane and electrode resistivities, electrolyte conductivity, gas bubble formation, and temperature on the total cell resistance. It integrates exergy-based metrics to quantify performance and provides optimization pathways to reduce ohmic overpotentials and enhance system efficiency [78]. Milewski et al. put forward two mathematical models of AWE; one of them emphasized focusing on factors such as calibration and other parameters associated with the electrolytic cell and the energy losses associated with it. Another model focuses on a reduced-order equivalent circuit and the internal electric resistance of the electrolyte [79]. Dhabi et al. studied the PV-electrolysis modeling by a DC/DC buck converter with MPPT control for an improved adaptation between the electrolysis and PV generator [80]. El-Askary et al. developed a numerical model to predict the H2 generation process in AWE. Their results indicated that reducing the main flow velocity, increasing current density, and minimizing the cathode-anode gap significantly enhanced hydrogen production efficiency, offering insights for optimizing electrolyzer design [81]. Tanaka et al. featured an NaCl electrolysis industry to produce chlorine and caustic soda, such as chlor-alkali, in an electrolyzer with a double-layer membrane consisting of sulfonic acid and carboxylic acid groups to prevent back migration of OH ions and succeeded in increasing voltage and decreasing current efficiency during electrolysis [82]. Shestakova et al. studied the ways to develop cost-effective and environment-friendly Ti/Ta2O3-SnO2 electrodes for H2O and organic compounds oxidation, and also characterized electrochemically and physically [83].
The future of AWE is deeply intertwined with advancements in high performance electrode materials, innovative membrane technologies, and the integration of renewable energy sources. The primary focus of research must be on developing cost-effective, durable, and highly active catalysts with optimized nanostructured surfaces to enhance both the HER and OER. By leveraging advanced nanomaterials, multi-metallic alloys, and engineered surface architectures, catalytic efficiency can be significantly improved, leading to lower overpotentials, reduced activation energy, and enhanced reaction kinetics. The incorporation of transition metal chalcogenides, phosphides, and carbides, alongside traditional Ni-based materials, is a promising direction for boosting catalytic performance while maintaining economic feasibility.
Beyond electrode optimization, membrane technology remains a critical area for improving ionic conductivity, chemical stability, and gas separation efficiency in AWE systems. Current diaphragm-based separators are prone to ionic resistance and degradation, necessitating the development of hybrid membranes, composite ion-exchange structures, and functionalized polymeric separators that offer long-term stability under alkaline conditions. These improvements will minimize crossover losses, enhance selectivity, and improve operational lifespan, directly impacting overall system efficiency. Additionally, high pressure electrolysis has gained attention as a means of reducing downstream compression costs while achieving higher volumetric hydrogen density. Investigating the long-term stability and material compatibility of electrodes and membranes in high pressure, high temperature alkaline environments will be crucial for industrial scale applications.
The integration of AWE with renewable energy sources, including solar photovoltaics, wind power, and nuclear energy, presents a significant opportunity for scaling hydrogen production while reducing carbon footprints. However, the intermittent nature of renewable energy poses challenges in maintaining stable electrolysis operation. Developing smart grid-compatible AWE systems with dynamic load management, real-time electrochemical monitoring, and adaptive power input regulation will enable better synchronization with fluctuating renewable energy outputs. Additionally, the hybridization of electrolysis technologies—such as combining AWE with proton exchange membrane (PEM) or solid oxide electrolysis (SOE) systems—could enhance operational flexibility, allowing for higher efficiency under varying power conditions. These hybrid approaches leverage the strengths of alkaline, protonic, and high temperature electrolysis to create more resilient and energy-efficient hydrogen production pathways.
Computational modeling and numerical simulations will continue to play a pivotal role in optimizing AWE systems. Multi-physics models integrating fluid dynamics, ion transport mechanisms, and electrochemical kinetics can provide deeper insights into system performance under various operational conditions. High-fidelity FEA and CFD simulations can aid in designing next-generation electrolyzers with improved flow dynamics, minimized bubble formation, and reduced mass transport losses. The incorporation of machine learning and artificial intelligence (AI)-driven predictive models will further refine process control strategies, enabling real-time efficiency improvements and fault detection mechanisms in industrial AWE applications.
Ultimately, positioning AWE as a cornerstone of a global hydrogen economy requires continued materials innovation, process optimization, and seamless integration with renewable energy infrastructure. By addressing fundamental challenges such as catalyst degradation, ionic resistance, and energy loss, AWE can evolve into a highly scalable, cost-effective, and environmentally sustainable hydrogen production technology. Future research should emphasize the convergence of experimental advancements with computational modeling to accelerate the development of next-generation high efficiency electrolyzers, ensuring economic feasibility and large-scale deployment in the transition toward a carbon-neutral energy landscape.

3.2. Proton Exchange Membrane (PEM) Electrolysis

Proton exchange membrane (PEM) electrolysis was developed in 1966 as an advanced alternative to AWE, addressing limitations such as low current densities, carbonate formation, low gas purity, and operational pressure constraints. In PEM electrolysis, a solid polymer electrolyte (SPE) serves multiple critical functions, including electrical insulation of electrodes, proton conduction, and efficient separation of product gases (Figure 6b). The fundamental electrode reactions, charge carrier transport, and overall cell operation in PEM electrolysis are detailed in Table 3.
PEM electrolysis offers several advantages over AWE, including lower gas permeability, high proton conductivity, compact design, and higher efficiency (80–90%), while producing high purity hydrogen with oxygen as a byproduct. However, its efficiency and long-term stability are influenced by various factors, such as current density, operating pressure, temperature, membrane type, and catalyst selection. Among these, current density plays a crucial role in performance and degradation rates. Rakousky et al. investigated the impact of constant and dynamic current density profiles, concluding that the highest degradation occurs at 2 A/cm2, while no significant degradation was observed at 1 A/cm2, highlighting the importance of optimized operating conditions for improving the longevity and reliability of PEM electrolyzers [84]. According to Zhou et al. [85], PEM-based systems benefit from higher current densities (typically 2–3 A/cm2), which enable compact designs and rapid dynamic responses. These features make PEM electrolysis particularly suitable for integration with intermittent renewable energy sources. Similarly, Wang et al. [86] reported that PEM electrolyzers can operate across a wide range of current densities, up to 3 A/cm2, while maintaining efficiency, due to advanced membrane and catalyst designs that minimize ohmic and activation losses. However, as Fahr et al. [87] pointed out, higher current densities also lead to challenges such as hydrogen crossover in PEM systems. Thinner membranes are desirable for reducing ohmic losses, but at high current densities, increased gas production can cause higher crossover rates, potentially leading to reduced efficiency and safety concerns if not properly managed. Rocha et al. [88] emphasized that gas bubble evacuation becomes critical at elevated current densities. Their research showed that in alkaline electrolysis systems, optimizing electrode topology can improve mass transport and bubble removal, significantly boosting performance. Their flow-engineered 3D electrodes allowed for PEM-like efficiency at high current densities using low-cost materials. Peng et al. [89] explored a novel integration of PEM electrolyzers with thermal energy storage systems to stabilize temperature fluctuations caused by varying power input. This helped maintain consistent performance even during frequent current density shifts—common when electrolyzers are powered by fluctuating renewables. Araújo et al. [90] further highlighted that high current densities, while beneficial for hydrogen output, exacerbate durability issues due to the stress imposed on noble-metal-based electrocatalysts. They argued for the development of PGM-minimized electrocatalysts that retain performance under such demanding conditions. Meanwhile, Wang et al. [91] compared PEM and AWE technologies, noting that AWE systems, while less tolerant of high current densities, still maintain a place in large-scale applications due to their lower capital cost. However, their lower operating current densities make them less flexible for variable power input scenarios.
Furthermore, studies have shown that catalyst composition significantly influences degradation in PEM electrolysis. Ito et al. investigated PEM using a platinum-based anode catalyst layer, revealing that at pressures below 10 bars, the permeated H2 flux was largely consumed through oxidation or recombination at the anode. These findings highlight the importance of optimized catalyst selection and operating pressure in mitigating hydrogen crossover and performance losses, ultimately enhancing the efficiency and durability of PEM electrolyzers. Grigoriev et al. described a numerical and an experimental analysis of the optimum loadings of platinum and iridium in PEMWE cells. It was reported that these metals content did not show substantial catalytic layer degradation within 4000 h. Based on the results obtained, the recommended optimal Pt and Ir loadings derived from the study are a Pt loading of about 0.4 mg/cm2 (or 1 mg/cm2 of Pt/C with 40 wt.% Pt) and an IrO2 loading of approximately 2.5 mg/cm2 [92]. However, costly Pt and Ir, presently considered as the state-of-the-art electro-catalysts, stop the extensive commercialization of this technology. Thus, Kus et al. [93] presented a cost-effective method for preparing Ir-based anode catalysts using TiC as a support. As shown in Figure 11a, their optimized TiC loading of 0.4 g/cm3 yielded the highest current density at the lowest cell voltage, demonstrating improved electrochemical performance. This integration of TiC significantly reduces noble metal loading while maintaining system efficiency. Rozain et al. further reported the applicability of titanium particles with IrO2 particles at the anode of PEMWE cells. According to MEA, for over 1000 h of operation, IrO2/Ti was found to be stable, but the degradation rate measured at 1 A cm−2 was reduced from 180 µV h−1 (for the pure IrO2 anode) to only 20 µV h−1 (for the 50 wt.% IrO2/Ti anode) [94]. Rozain et al. have studied PEMWE using IrO2 alone as a catalyst at 80 °C. They used 0.25 mg/cm2 of Pt cathode and 0.5 mg/cm2 of IrO2 anode at 1 A/cm2 of current density with a Nafion 115 membrane [95]. Lee et al. have also developed electrodeposited IrO2 anodes in PEMWE. At 0.7 V, the IrO2 loading ranges from 0.007 to 0.464 mg/cm2 at 90 °C in the PEMWE test. An IrO2 loading of 0.1 mg/cm2 with the IrO2/CP electrode shows the highest performance [96]. Kadakia et al. investigated a nanostructured fluorine (F)-doped IrO2 electrocatalyst for PEMWE, exploring compositions ranging from 0–20 wt.% F content. Their results demonstrated that the electrocatalyst follows a two-electron transfer reaction mechanism, with an activation energy of approximately 25 kJ/mol for electrolysis. Notably, the 10 wt.% F-doped IrO2 composition exhibited the highest electrochemical activity, which was further validated through single full cell tests, confirming its potential for enhancing PEMWE performance and catalytic efficiency [97]. Lamy et al. used PtSnRu/C, PtSn/C, and Pt/C catalysts to yield hydrogen by electrolysis and studied the ethanol’s electrocatalytic oxidation in a PEMWE cell at 20 °C. It was observed that the voltage did not exceed 0.9 V at 100 mA/cm2 for a 220 cm3 evolution rate of H2 per hour. Also, the consumption of electrical energy was less than 2.3 kWh, which was one-half of the energy required for the water electrolysis [98]. Kadakia et al. also stated that a nanostructured SnO2 with 10 wt.% or 20 wt.% RuO2 improved the durability and the electrochemical activity as compared to noble metal counterparts [99]. Sapountzi et al. discussed the effect of the structure and nature of the catalyst-electrode materials performance, whether it be AWE, PEMWE, or SOEC [100]. Zlotorowicz et al. fabricated zirconium hydrogen phosphate, Nafion, and iridium oxide on glassy carbon disk electrodes. At room temperature, the electrocatalyst was electrochemically examined with OER in 0.5 M sulfuric acid electrolyte. The outcomes concluded that the existence of zirconium hydrogen phosphate worsens the current per mass and emphasized the need to improve the catalytic layers [101]. Figure 11b shows the curve between mass activity and anode loading. It is clear that increasing anode loading mass activity decreases. Figure 11c shows the curve between efficiency and time of 0.10 mg/cm2 of pure IrO2 and 0.12 mg/cm2 of IrO2/Ti particles. It is clear that IrO2/Ti shows better efficiency as compared to pure IrO2. Figure 11d shows the curve of how efficiency is affected by different IrO2 loadings. It is clear from the plot that 0.5 mg/cm2 of IrO2 shows the maximum efficiency of 71%. Figure 11e displays the curve between the cell voltage and the current density when the IrO2 membrane was doped with fluorine, and Figure 11f demonstrates the curve between the cell voltage and the current density when RuO2 was used with IrO2.
As the name itself suggests, the membrane also provides a major influence on the efficiency of a PEMWE cell. Rakousky et al. demonstrated that the membrane plays a major role in determining the overall efficiency of a PEMWE cell. Although structurally stable, its performance is influenced by contamination from titanium species and changes in ion conductivity. Their findings emphasize that membrane integrity, along with compatible cell components, is critical for minimizing degradation and improving efficiency [102]. Skulimowska et al. in their work, observed that perfluorosulfonic acid Aquivion ionomer has given higher water electrolysis performance as compared to the Aquivion zirconium phosphate membrane [103].
Wang et al. used a co-crystallized catalyst-coated membrane (CCM), prepared by heating the catalyst layer and amorphous Nafion membrane at 120 °C, and used the solid polymer as the electrolyte. According to results, a decrease in cell voltage by 0.009 V at 2000 mA/cm2 at 80 °C with improved stability was noticed as compared to the conventional CCM [104].
Moving further, variation in the efficiency of a PEMWE with the varying pressure is also reported. Olesen et al. have used a circular planar, inter-digitated flow field at the anode for a high pressure PEMWE cell [105]. At the differential and balanced pressure operation, Schalenbach et al. compared the in-situ measurements of the anodic H2 content, and simulated the effect of cathodic and anodic pressures in PEMWE on the gas crossover [106]. Lamy discussed both low temperature PEMWE cells and high temperature SOEC cells [107]. Olivier et al. studied the modeling process for the coupling of the PEM electrolysis to intermittent electrical sources. The model was proved to accurately predict the dynamic behavior of a semi-industrial PEM electrolysis system [108]. Langemann et al. have discussed the bipolar plate as a multi-functional component during PEMWE and tested Au and TiN coatings in PEMWE environments for the applications as a protective layer [109]. Millet provided a summary of PEMWE, including the effects of operating temperature and pressure on voltage and the structure of PEMWE [110]. Hancke et al. investigated the performance of a PEM water electrolyzer under pressures up to 180 bar, revealing that efficiency improves slightly up to 30 bar but declines significantly at higher pressures. The study highlights increased hydrogen crossover and ohmic losses as key factors behind reduced efficiency [111]. Sun et al. developed a 9-cell PEMWE stack and examined it for 7800 h. As observed, Fe, Cu, and Ca were distributed in the membrane and the catalyst layer of the CCMs. The cathode and the anode overpotential increased because the cations reside in the Nafion polymer electrolyte in the membrane and the catalyst layer [112]. Yilmaz et al. drove PEMWE by geothermal power for H2 production at 160 °C. The 3810-kW of power produced in a binary geothermal power plant is for the electrolysis process. It was reported that the preheated water (by the geothermal water) used in electrolysis produced H2 at a rate of 0.0340 kg/s. The exergy and the energy efficiency were 45.1% and 11.4%, respectively, of the binary geothermal power plant [113]. Rozain et al. reported the electrochemical characterization of PEMWE. According to results, at voltages 1.8–1.9 V, the charge transfer processes were the main cell impedance contributors. Also, the impedance was insignificant with the HER and two time constants on experimental impedance spectra with the OER [114]. Bensmann et al. discussed an in-situ method for the determination of a fully assembled cell by the use of standard equipment. The method was briefly illustrated for a broad pressure range with a laboratory-scale electrolyzer, and the measured data were compared with available literature values [115].
PEM electrolysis is a high efficiency hydrogen production technology, offering compact design, high proton conductivity, and gas purity, with efficiencies reaching 80–90%. However, commercialization is limited by the high cost of platinum group metal (PGM) catalysts, membrane degradation, and operational stability concerns. Iridium-based catalysts, such as IrO2 and IrO2/Ti, exhibit exceptional long-term stability, with IrO2/Ti electrodes maintaining performance beyond 1000 h while demonstrating lower degradation rates. Advancements in nanostructured and doped catalysts, including TiC-supported iridium, fluorine-doped IrO2, and RuO2-based electrocatalysts, have shown improved electrochemical performance and reduced overpotential losses, presenting viable alternatives to reduce material costs. Furthermore, non-PGM catalysts, such as transition metal oxides and carbon-based electrocatalysts, are being explored to enhance catalytic activity while improving cost-effectiveness.
Beyond catalysts, membrane stability remains a crucial challenge, as ionic resistance and gas crossover contribute to performance degradation. Studies indicate that advanced perfluorosulfonic acid (PFSA)-based ionomers, such as short-sidechain Aquivion®, can outperform conventional Nafion membranes in PEMWE applications. These alternative PFSAs offer higher proton conductivity and improved chemical stability, particularly under high current density and temperature conditions, making them attractive candidates for durable and efficient membrane formulations. Porous transport layers (PTLs) play a significant role in PEMWE performance, where Pt-coated Ti-PTLs have been shown to reduce degradation rates and enhance electrode durability compared to conventional PTLs. High pressure PEM electrolysis has enabled direct hydrogen production at elevated pressures, reducing downstream compression costs, though long-term stability studies remain limited.
Operational parameters, including temperature, pressure, and current density, also play a crucial role in PEMWE efficiency (Figure 12). The Figure 12 illustrates how temperature, pressure, membrane thickness, and current density affect overall system efficiency. Efficiency increases with temperature due to enhanced reaction kinetics and lower ohmic resistance, while higher current densities reduce efficiency due to elevated overpotentials. Membrane thickness influences ohmic losses, and pressure impacts thermodynamic potential. Higher temperatures improve reaction kinetics and electrode activity, but excessive thermal exposure can accelerate degradation rates.
Computational modeling and FEA have been instrumental in optimizing cell architecture, charge transport, and reaction kinetics, contributing to next-generation high efficiency electrolyzers. Hybrid approaches, such as co-crystallized catalyst-coated membranes (CCMs) and renewable energy integration, are further improving system efficiency and sustainability.
Despite these advancements, economic feasibility remains a significant challenge, necessitating progress in scalable manufacturing, cost-effective catalysts, and energy management strategies. Future research should focus on long-term durability assessments exceeding 10,000 operational hours, enhanced membrane conductivity, and real-time electrochemical monitoring to ensure PEM electrolysis remains a cornerstone of the hydrogen economy, driving cost-effective and sustainable large-scale hydrogen production.

3.3. Solid Oxide Electrolysis Cell (SOEC)

SOEC is first reported in 1980 and runs in regenerative mode for the electrolysis of water to produce oxygen gas, as shown in Figure 6c. SOEC has attracted attention because it converts electrical energy to chemical energy for ultra-pure H2 with 90–100% efficiency. The significant characteristics of SOEC are its working conditions, i.e., high pressure and temperatures. Also, it uses the water in steam form and O2− conductors as a charge carrier, as presented in Table 3.
Lim et al. have investigated the effects of seawater electrolysis by electrolyzing steam from a simulated seawater bath by Ni-YSZ/YSZ/LSCF-GDC SOEC for H2 production. It was observed that the sea salt impregnation vaporized at 800 °C during the operation period [116]. In a study by Deka et al., the electrochemical behavior of a nickel-doped, A-site deficient lanthanum strontium ferrite (La0.7Sr0.2FeO3) was evaluated as an SOEC cathode material. Electrolysis was performed at 800 °C with a 3% H2O/He gas stream across various current densities. The study revealed that the undoped La0.7Sr0.2FeO3 exhibited reduced Faradaic efficiency for hydrogen production due to the formation of a non-conductive La2O3 secondary phase, which impaired its electrochemical performance. In contrast, the nickel-doped version achieved nearly 100% Faradaic efficiency. X-ray diffraction (XRD) analysis showed that nickel doping promoted the formation of a conductive B-site metal phase and a Ruddlesden–Popper phase with mixed ionic and electronic conductivity, significantly enhancing the cathode’s performance [117]. Moreover, Mizusawa et al. investigated the temperature distribution during high temperature steam electrolysis in a 2-D tubular model of a micro-tubular SOEC. The results displayed that the current collecting positions were strongly affected by the temperature and reaction distributions [118]. The use of Ce0.6Mn0.3Fe0.1O2− (CMF) as a new oxide cathode in SOEC using LaGaO3-based electrolyte for H2 production at high temperature up to 1173–973 K was observed in an experiment by Hosoi et al. and it was observed that CMF could be a proficient oxide cathode for intermediate temperature SOEC [119]. Zhang et al. showed the 4 kW HTSE long-term test results, completed at Idaho National Laboratory (INL). The results concluded that a 3.1% degradation rate in 830 h of stable operation was achieved at a current density of 0.41 A/cm2 [120]. Reytier et al. carried out experiments in both electrolysis and co-electrolysis in an HTSE based on SOEC. A 10- and 25-cell stack tested and produced 0.6 and 1.7 Nm3/h of H2 at 800 °C, respectively, in the electrolysis mode below 1.3 V [121]. In an experimental study by Alenazey et al., they focused the production of CO and H2 by SOEC and observed that the production of H2O and CO2 increased by enhancing the current density [122]. Cacciuttolo et al. tested the effect of pressure on HTSE based on SOEC; a positive effect of pressure on the oxygen side performances was reported [123]. Kazempoor et al. have produced synthetic gas using HTSE based on SOEC and reported a numerical model. The model has adopted a moderate fidelity approach to predict the thermo-fluid and electrochemical phenomena inside the cell and showed that the operating temperature, flow rates, and fuel consumption have a direct effect on power consumption and SOEC performance [124]. Ardigo et al. studied the K41X stainless steel for HTSE based on SOEC. It was observed that the alloy displayed an excellent oxidation resistance as compared to single atmosphere tests [125]. In an experiment, Houaijia et al. developed a solar hydrogen high temperature electrolysis to superheat steam up to 700 °C. The receiver was operated in a DLR’d solar simulator of steam reaching about 700 °C at a 40% thermal efficiency, a 4 kW solar power, and at a thermal-to-hydrogen efficiency of 26% [126]. Mougin et al. described the durability and performance in the stack environment in SOEC. It was shown that for SOFC, the protective coatings were compulsory to reduce the degradation rate in HTSE stacks [127]. Giddey et al. demonstrated SOEC at near room temperature and investigated the thermodynamic and practical energy benefits of a single-step water electrolysis process assisted by carbon [128].
The variation in electrodes (both anode and cathode), or cathode alone, or anode alone causes variation in the degradation rate. To improve the gas diffusion, Dong et al. prepared micro-channeled cathode support by a mesh-templating phase-inversion process. It was observed that this cathode support increased the H2 production and steam utilization efficiency [129]. Hauch et al. used Ni/YSZ microstructure electrodes in SOEC and made a fine and denser NiO/YSZ precursor electrode for extended stability of electrolysis at high p(H2O) and high current densities. Applying such conditions, they showed a 0.3–0.4%/kh degradation rate of 1 A/cm2 at 800 °C [130]. Mahmood et al. developed a YSZ electrolyte membrane for a solid electrolyte membrane reactor and used electrochemical activity of CO2, steam, and a CO2-steam mixture at 800 °C. A very high current density was detected for CO2, steam, and CO2-steam mixture electrolysis at 850 °C and 1.5 V [131]. Hou et al. used an MoO2-based cathode in their experiment of co-electrolysis based on SOEC at 750 °C. It was observed that, the MoO2-based and the Ni/YSZ cells showed alike electrochemical performance in the co-electrolysis mode, while in the CO2 electrolysis mode, the MoO2 based cell displayed a better electrocatalyst as compared to the Ni/YSZ cell [132]. Nechache et al. studied the behavior and electrochemical performance of a commercial electrode-supported cell of Ni-YSZ/YSZ/LSCF [133]. Further, Li et al. have used a composite La0.8Sr0.2MnO3− (LSM) in SOEC and reported the maximum current efficiency of 65% of Fe2O3-loaded LSM composite in a proton-conducting solid oxide electrolyzer at 800 °C [134]. Xing et al. performed co-electrolysis using an LSM-based electrode and observed lower ohmic resistance and better electrochemical performance [135]. Chen et al. doped scandium into LSCM to improve the performance of the composite cathode. The doping improved the ionic conductivity but decreased the mixed conductivity. Moreover, faradic efficiencies were also improved [136]. Li et al. used a composite-based LSCM cathode and used that for direct carbon dioxide electrolysis with copper nanoparticles at the surface of the LSCM cathode. It was observed that 85% current efficiencies were obtained with an LSCM cathode for direct CO2 electrolysis in an oxide-ion-conducting SOE [137]. Ge et al. introduced Sr2FeNbO6 (SFN) to the SOEC as the H2 electrode. Higher conductivity was observed in the H2/H2O atmosphere as compared to that in the air. Also, if compared with Ni/YSZ, SFN-YSZ was more appropriate for the H2 electrode in SOEC. Ge et al. introduced Sr2FeNbO6 (SFN) as a hydrogen electrode material in SOEC, reporting high electronic conductivity (2.215 S·cm−1 at 850 °C in H2/H2O), which surpasses performance in air and supports its suitability for steam electrolysis environments. Compared to conventional Ni/YSZ, the SFN-YSZ composite showed reduced charge transfer resistance and shifted the rate-determining step from interfacial charge transfer to surface adsorption/diffusion—indicating more favorable catalytic behavior. Ni/YSZ, while widely adopted, is prone to degradation via Ni sintering and carbon deposition, especially under CO-rich or hydrocarbon conditions [138]. Although Li et al. showed carbon deposition on Ni/YSZ could be promoted in SOEC mode, such effects are typically detrimental for long-term operation due to increased overpotential and material degradation [139]. Keane et al. have also analyzed the role of micro-structured Ni-YSZ cathodes on degradation rate in SOECs and observed significant electrochemical degradation in cells operated at 840 °C [140]. Ruan et al. performed CO2 electrolysis with a composite cathode with ceramic, iron nano-catalyst, and iron oxide catalyst loading on it. The loading of nano-catalyst improved the electrode performance, and the current efficiency of CO2 electrolysis was enhanced by 5% for the LSCM electrode at 800 °C [141]. Yoon et al. used lanthanum strontium vanadate (LSV) as the cathode for SOEC for co-electrolysis of steam [142]. Dong et al. have also reported A-site deficient and B-site excess perovskite LSCNNi composite cathodes. The results signified that the Ni nanoparticles cathode was a potential candidate for direct steam electrolysis in SOEC [143]. Xu et al. studied the Fe catalyst over the composite electrode for direct steam electrolysis. It was observed that the Fe catalyst enhanced the faradic efficiency and the electrode performance without the flow of reducing gas over the cathode. The current efficacy was also improved by 30% and 40% as compared to the bare LSCM-based cathodes at 800 °C [144]. Dasari et al. reported the electrochemical characterization of the Ni/YSZ electrode for H2 production in SOEC [145]. Li et al. further loaded LSCM electrodes with Ni for SOEC, and the faradic efficiency was enhanced by 20% as compared to the bare LSCM cathodes [146]. Yang et al. used a K2NiF4-type structured Pr0.8Sr1.2(CoFe)0.8Nb0.2O4+ (K-PSCFN) matrix with a nano-sized Co-Fe alloy (CFA) electrode with LSGM electrolyte in SOEC at 900 °C. The cell confirmed good stability for high temperature steam electrolysis and an H2 production rate calculated from Faraday’s law under a voltage of 1.3 V at 900 °C [147]. Liu et al. used an asymmetric NiO-YSZ cathode substrate made up by the phase-inversion tape casting method. The electrochemical performance was enhanced and displayed an excellent current density with a high production rate at 800 °C [148]. Li et al. employed an infiltration method on the composite electrodes in order to accomplish an activity-enhanced electrode performance. The partial pressure of 5–100% was achieved for Fe-loaded LSV cathode in a wide range of H2 [149]. Gan et al. observed that the electro-catalytic activity of La0.4Sr0.4TiO3 (LSTO) was insufficient for efficient electrochemical reduction of steam or CO2. To enhance performance, catalytically active Ni nanoparticles were incorporated into the LSTO cathode via an impregnation method, significantly improving its activity for direct steam electrolysis. As a result, the current efficiency was notably enhanced compared to the bare LSTO-SDC cathode, achieving superior performance under 2.0 V at 800 °C [150]. Ardigo et al. tested La0.8Sr0.2MnO2− and LaNi0.6Fe0.4O2− as coatings in an O2-H2O atmosphere at 800 °C. The results concluded that LaNi0.6Fe0.4O2− coating has shown high temperature corrosion resistance and enhanced electrical conductivity [151].
The electrolyte plays a vital role in varying the degradation rate of electrolysis. Sumi et al. have used BaCe0.8Y0.2O3− (BCY) between Ce0.9Gd0.1O1.95 (GDC) and the Ni-GDC fuel electrode. The leakage of current by the ceria-based electrolyte raised in SOEC because electron conductivity appears at low O2 partial pressures, and BCY stopped the current leakage. Also, the performance of SOEC was superior than a conventional electrolyte (YSZ) by using the BCY blocking layers and the GDC electrolyte at 500 °C [152]. Pu et al. used BaCe0.8Y0.2O3− in SOEC at 600 °C, which was lower than the conventional SOEC composed of YSZ (generally above 800 °C). It was observed that the performance electrolysis was enhanced as compared to conventional [152]. Further, Hjalmarsson et al. have reported SOEC using Ni/YSZ cermet support electrodes with a YSZ electrolyte. The cell was tested at 800 °C and −1 A/cm2 converting 31% of a combination of H2:H2O:CO2 for 2700 h. The results showed electrode degradation in 350 h followed by partial reactivation [153]. Lauka et al. have used wood ash as a solid electrocatalyst, similar to zeolite structure. The results showed a strong correlation between the used sample mixture and pH value [154]. Farandos et al. established stable aqueous colloidal dispersion of YSZ and utilized them to construct 2D planar and 3D microstructures by inkjet printing [155]. Zhang et al. discussed the dynamic state effect of the low concentration of Na+ in the SPE water electrolyzer. It was observed that in the presence of Na+, the cell performance degraded more severely [156]. Wang et al. fabricated a CCM for SPE water electrolysis by partially crystallizing a Nafion membrane and catalyst layers. The results revealed that the electrolysis voltage of the SPE water electrolysis with the new CCM was low at 80 °C and atmospheric pressure [157].
The H2O and CO2 co-electrolysis in an SOEC is capable of energy storage and utilization of CO2. Xu et al. numerically studied the effects of CH4 on CO2/H2O co-electrolysis by a 2D model. It was observed that the CH4 support in dropping the equilibrium SOEC potential hence the electrical power consumption reduces drastically [158]. Tao et al. have reported Ni-YSZ-based SOEC co-electrolysis at a current density of −1.5 or −2.0 A/cm2. The thorough electrochemical analysis discovered that there was a substantial increase in the oxide ion transport resistance [159]. Luo et al. established a 2D dynamic model for the dynamic response of CO2/H2O co-electrolysis in tubular SOEC (TSOEC) [160]. Menon et al. investigated the behavior of an SOEC co-electrolysis by the interactions between electrochemical parameters and transport processes. The influence of microstructural properties, temperature, and cathode gas velocity are discussed [161]. Stempien et al. reported a comparative study on the equilibrium potential of H2O and CO2 SOEC co-electrolysis. It was concluded that at temperatures above 800 °C, when the concentration of CO2 is below 25%, then the oxygen partial pressure model suffices for modern cells [162]. Luo et al. analyzed the efficiency and the performance of H2O/CO2 co-electrolysis in TSOEC. The results showed that the CO2 conversion ratio was significantly promoted by the reversed water gas shift reaction [163]. Li et al. have tested the H2O-CO2 co-electrolysis performance and mechanism of SOEC at 550–750 °C operating temperatures and found out that the co-electrolysis performance significantly increased with respect to the temperature for the Ni/YSZ/ScSZ/LSM-ScSZ electrolysis cell. Also, the CH4 production is promoted by electricity and effectively suppressed by Ru in porous Ni/SYZ cathode [164]. In an experiment, Kyriakou et al. produced hydrogen from steam electrolysis and simultaneously converted methane to hydrocarbons. The reaction system was investigated in a solid-state oxygen ion (O2−) conducting cell at temperatures between 700 °C and 840 °C, and it was also observed that the conversion process was enhanced with the use of SZY on the anodic electrode to serve as a methane coupling catalyst [165]. Patcharavorachot et al. investigated the impact of methane presence on the performance of solid oxide electrolysis cells (SOECs) and introduced a Solid Oxide Fuel-Assisted Electrochemical Cell (SOFAEC) model to evaluate its efficiency. Their study analyzed energy efficiency and power input under conditions with and without CH4. The findings revealed that SOFAEC exhibited superior performance compared to conventional SOECs, demonstrating enhanced electrochemical efficiency and reduced energy consumption, highlighting the potential of methane-assisted electrolysis in improving hydrogen production efficiency [166]. Aicart et al. performed co-electrolysis at 800 °C and reported the relevance of the macroscopic representation of electrochemical processes through a “surface ratio” that takes into account the co-electrolysis [167]. Li et al. reported the electrochemical reactions, the heterogeneous elementary reactions, the transport of mass and charge, and the electrode microstructure in a 1D elementary reaction of H2O/CO2 co-electrolysis in SOEC. The simulation results concluded that the heterogeneous reactions occur near the cathode, and the electrochemical reactions occur in the electrode [168].
The rate of degradation is affected by the variation in current density, resistance, or voltage. Zheng et al. investigated the air electrode contact, H2 electrode contact, and the electrochemical contributions of the cell at 750 °C in the SOEC stack with a 90/10 H2O/H2 ratio. At a constant current density, the air electrode contact, H2 electrode contact, and the voltage degradation of the cell were 19.6%, 8.9%, and 71.5%, respectively, of the total voltage degradation [169]. Shimada et al. [170] demonstrated that achieving current densities over 3 A/cm2 is feasible using nanocomposite oxygen electrodes fabricated via spray pyrolysis. The bimodal-structured electrodes composed of Sm0.5Sr0.5CoO3−δ and Ce0.8Sm0.2O1.9 enabled high performance due to improved conductivity and gas diffusion. These cells achieved 3.13 A/cm2 at 750 °C and 4.08 A/cm2 at 800 °C, directly correlating high current densities with increased hydrogen production rates. Similarly, Kim et al. [171] emphasized that the electrode microstructure plays a critical role under high current density operations. By tailoring the air electrode using graphite pore formers, they improved both the electrochemical performance and long-term durability. The enhanced porosity and triple-phase boundary density helped maintain performance even under demanding fuel cell–electrolyzer cycles. Schefold et al. [172] conducted long-term testing (over 80,000 ON/OFF current cycles) on SOECs and found that steady operation at a current density of −0.7 A/cm2 resulted in minimal degradation—only 5 mV/1000 h. This underscores the feasibility of stable operation under moderate current densities, especially relevant for renewable energy integration where intermittent loading is common. Gaikwad et al. [173] reviewed that while SOECs offer the highest hydrogen generation efficiency among electrolysis technologies, Faradaic efficiency declines at high current densities if gas diffusion and electrode-electrolyte interfaces are not optimized. High current densities increase reaction rates, which can outpace the material’s ability to manage ionic and electronic transport effectively. Jing et al. [174] elaborated on the transition from O2−- to H+-conducting electrolytes in SOECs, especially under intermediate temperatures (400–700 °C). At higher current densities, proton-conducting SOECs (H+-SOECs) offer the advantage of lower activation energies and higher ionic conductivity, making them better suited for high rate operation with improved electrical efficiency. Laguna-Bercero [175] highlighted in high temperature electrolysis that SOECs can achieve efficiencies above 80% (LHV) at high current densities when heat and steam from renewable sources are integrated. Xu et al. [176] investigated the dynamic response of SOECs under fluctuating current densities, particularly in the context of renewable energy integration. They emphasized that rapid shifts in current density led to thermal gradients and mechanical stress, particularly at the electrode-electrolyte interface, which can cause cracking and performance decay. Their study found that maintaining temperature gradients below 10 K/cm is critical for structural safety. Additionally, they highlighted the need for control strategies to manage real-time variation in current density caused by unstable renewable energy sources such as wind or PV to prevent thermal fatigue and irreversible damage. Wolf et al. [177] addressed the industrial implications of operating SOECs at high current densities. Their review of commercial systems, including projects such as GrInHy2.0, revealed that current densities around −0.9 A cm−2 are achievable over long durations (up to 23,000 h) with acceptable degradation rates (e.g., 0.5–1.5% per 1000 h). They pointed out that despite SOECs achieving near 100% electrical efficiency under ideal conditions, the long-term viability at high current densities demands tailored materials and thermal integration with upstream or downstream processes (e.g., using waste heat from steel production).
Tanaka et al. developed a quasi-1D simulation model for overvoltage and cell voltage in high temperature SOEC. Results showed that there was an improved accuracy against cell voltages at 750–850 °C. Furthermore, the case studies at 800 °C discovered that at 1.0 A/cm2 and 82% steam utilization, 40% of local current density distribution in the cell [178]. Bermeio et al. analyzed different thermodynamic cell operational modes and operational strategies. The study revealed that the H2 production system achieved a flat performance curve within the complete power load range, with overall efficiencies between 91 and 97% as compared to HHV [179]. Peters et al. also investigated the operating conditions and the system configurations to study the efficiencies of producing H2 by SOEC. The calculated efficiencies vary from 104% to 62% on the basis of lower heating value of H2 produced [180]. Zheng et al. investigated the NiO/YSZ/GDC/LSCF-GDC SOEC stack at the H2O/H2 ratios of 70/30, 80/20, and 90/10 at 750 °C. The investigations indicated that the cell resistance was 76.3–66.7% that of the stack repeating unit (SRU) and 23.6–27% of the contact resistance between the air electrode current collecting layer and the interconnect account [181]. Petrakopoulou et al. studied the four hybrid systems and the four different structures of a steam electrolysis system for H2 production. As observed, the maximum efficiency was achieved with a recycling sweep gas stream, which was further utilized in the border [182]. Kasai clarified the effectiveness of H2 for energy storage by high temperature steam electrolysis as compared to power storage by solar energy and nuclear energy [183].
The operating conditions, experimental setup or process of analysis have been discussed to optimize the electrochemical performance of SOEC. Cumming et al. have used a thermal camera to detect changes in the cell temperature by a remote, non-contact, and highly sensitive method [184]. Butler et al. examined the effect of the heat utilization in SOEC efficiency and H2-specific cell area on a 1D electrochemical model. A sensitivity analysis indicated the increased cost, the higher life, and the decreased power-to-gas storage of the SOEC stack [185]. Wu et al. developed a solid oxide direct carbon-assisted steam electrolysis cell in both the carbon bed and the cell, coupling at 800 °C. After analyzing, it was indicated that the system performance and production rates of gaseous fuels were highly influenced by working voltages of the cell and the carbon bed [186]. Henke et al. performed a theoretical study on the pressurized operation of SOEC in a range between 0.05 and 2 MPa. It was shown that at a low current density and low pressure, the electrolysis cell displayed better performance, and it improved with pressure at high densities [187].
Elder et al. operated SOEC between 500 °C and 900 °C to reduce CO2 to CO. It was stated that operating at the high temperature gave much higher efficiencies than low-operating electrolysis. Also, the electrode must be both electron and oxide ion-conducting, and minimizing the active surface area is necessary for efficient operation [188]. Mougin has focused on hydrogen production in SOEC and provided an overview of the advantages and challenges of this technology compared with other electrolysis technologies such as AWE or PEMWE [189]. Nechache et al. demonstrated how electrochemical impedance spectroscopy (EIS) is used to characterize the mechanism and the performance of SOEC electrodes and also as a corresponding instrument to study SOEC degradation processes [190]. Luo et al. developed a 1D elementary reaction kinetic model for SOEC with electrode microstructure, electrochemical reaction kinetics, heterogeneous elementary reactions, and transport processes of charge and mass. It was shown that SOFEC saved 80% of electricity at 3000 A/m2 and SOFEC displayed better performance than SOEC [191]. Mocoteguy et al. reviewed the cell degradation phenomena under high temperature electrolysis [192]. Jiang et al. reported the electrolysis system and seawater desalination on the basis of the newly invented triboelectric nanogenerator (TENG) [193].
The performance of SOECs heavily depends on the choice of materials for electrodes and electrolytes (Table 4 and Table 5). YSZ remains the conventional electrolyte due to its stability at high temperatures (800–1000 °C). However, alternative electrolytes such as gadolinium-doped ceria (GDC) and barium cerate (BCY) have shown promise in reducing operating temperatures while maintaining high ionic conductivity. The development of advanced cathode materials, such as LSV and Sr2FeNbO6 (SFN), has improved conductivity and efficiency, reducing degradation over long-term operation.
Solid oxide electrolysis cells (SOECs) have demonstrated significant potential for high efficiency hydrogen production, achieving 90–100% efficiency in high temperature steam electrolysis (HTSE). However, material degradation, system stability, and cost-effectiveness remain key barriers to large-scale commercialization. The primary degradation mechanisms include electrode delamination, phase segregation, and microstructural instability at elevated temperatures. Studies have reported degradation rates ranging from 0.3% to 3.1% per 1000 h, influenced by current density, operating temperature, and electrode-electrolyte interactions. Zhang et al. observed a 3.1% degradation over 830 h at 0.41 A/cm2, while Hauch et al. optimized Ni-YSZ microstructures, reducing degradation to 0.3% under similar conditions. Despite its advantages, SOEC performance and stability are highly dependent on electrode and electrolyte selection (Table 4 and Table 5). YSZ-based electrolytes remain the standard due to their stability at 800–1000 °C, but emerging materials such as BCY, GDC, and LSGM offer lower operating temperatures with improved ionic conductivity. Cathode materials, including LSCM, LSV, SFN, and MoO2, have been explored to improve electrochemical activity and durability, with SFN demonstrating higher conductivity and LSCM showing superior efficiency in direct steam electrolysis. Operational parameters, particularly temperature and gas composition, significantly impact performance. Higher temperatures (>900 °C) improve reaction kinetics, as Elder et al. demonstrated, but also accelerate degradation. Studies on co-electrolysis of H2O and CO2 have shown promising results in synthetic fuel production and energy storage, with methane presence reducing equilibrium potential and power consumption. Future research should focus on developing advanced electrode materials, such as perovskite-based catalysts and infiltration-modified composites, to enhance catalytic activity and long-term stability. Additionally, hybrid SOEC configurations, integrating renewable energy sources such as solar and wind, could create a sustainable hydrogen production pathway. Economic feasibility will depend on reducing material costs, optimizing manufacturing processes, and scaling up production. Advances in computational modeling and real-time electrochemical monitoring will further drive system efficiency improvements. While challenges remain, continued research in SOEC technology can position it as a key enabler of a hydrogen-based energy economy, providing an efficient and scalable solution for future storage of renewable energy derived electricity.

3.4. Microbial Electrolysis Cell (MEC)

Microbial Electrolysis Cells (MECs), first introduced in 2005, share fundamental principles with Microbial Fuel Cells (MFCs) but operate under opposite electrochemical conditions. Unlike MFCs, which generate electricity from organic matter, MECs convert electrical energy into chemical energy to produce hydrogen (H2) from organic substrates by applying an external voltage (Figure 6d). The anodic reactions involve microbial oxidation of the substrate, resulting in the release of H+, electrons, and CO2. The electrochemical potential generated at the anode facilitates proton migration through the electrolyte toward the cathode, where they combine with electrons to form hydrogen gas. The fundamental reactions, charge carriers, and overall cell mechanisms for MEC electrolysis are outlined in Table 3.
One of the key advantages of MECs is their ability to utilize wastewater as a substrate, providing a sustainable approach to hydrogen production while simultaneously treating organic waste. However, the technology remains limited to laboratory-scale research due to low H2 production rates, hydrogen purity challenges, high internal resistance, costly electrode materials, and complex system designs. High ohmic losses and biofilm limitations further restrict large-scale application. Over the past decade, extensive research has focused on optimizing electrode materials, reducing internal resistance, and improving microbial community efficiency to enhance overall system performance. Future advancements in catalyst development, system integration, and hybrid MEC configurations could pave the way for commercial viability, positioning MECs as a potential solution for sustainable hydrogen production and wastewater treatment within the water-energy-resource nexus.
A growing body of research has focused on enhancing electrode materials, improving bioanode efficiency, and optimizing cathodic reactions to overcome the current limitations of MEC technology and enhance hydrogen production yields. Ji et al. [194] investigated the use of Fe2+-modified biochar electrodes in MECs for phosphorus recovery from wastewater. Their study demonstrated a significant increase in current density and phosphorus removal efficiency, with power consumption reduced to 0.25 ± 0.01 kWh/kg P. This phosphorus-enriched biochar was also found to enhance soil quality when applied as a fertilizer. Samsudeen et al. [195] explored hydrogen generation using MECs integrated with anaerobic digesters for distillery wastewater treatment. Their modified MEC design yielded a maximum hydrogen production of 30.2 ± 1.2 mL at a current density of 811.33 ± 20 mA/m2. The compact retrofit design also showed a substantial reduction in chemical oxygen demand (COD) and improved energy recovery, demonstrating the feasibility of MECs for decentralized applications. Seelajaroen et al. [196] advanced MEC performance by modifying electrodes with poly(neutral red) and chitosan. Their findings highlighted enhanced COD removal (up to 67%) and methane yield (0.14 L CH4/gCOD), showcasing the dual potential of MECs in both carbon removal and biogas production. Chung et al. [197] focused on microbial hydrogen peroxide-producing cells (MPPCs), a subclass of MECs designed for H2O2 synthesis via the cathodic oxygen reduction reaction. Their review emphasized the environmental applications of bio-electrochemically generated H2O2, especially for water disinfection and pollutant degradation. Kadier et al. [198] optimized the operation of MECs for palm oil mill effluent (POME) treatment. Using response surface methodology (RSM), the study identified optimal conditions for maximizing hydrogen production and COD removal, highlighting the applicability of statistical modeling in scaling up MECs. Gonzalez et al. [199] examined various wastewater management technologies, including MECs. It emphasized the role of MECs in integrating with anaerobic membrane bioreactors and nature-based systems, especially in regions lacking centralized wastewater infrastructure. Cui and Yin [200] provided a comprehensive review of MEC applications for treating acid mine drainage (AMD), coupling pollutant removal with renewable hydrogen production. Their work stressed the effectiveness of MECs in acidic conditions and the potential for integrating with other technologies such as chemical precipitation and membrane filtration. Murugaiyan et al. [201] presented a detailed analysis of MEC configurations, including single and dual-chamber, packed-bed, and fluidized-bed systems. They concluded that while MECs are effective for simultaneous biohydrogen production and wastewater treatment, challenges such as high membrane and electrode costs remain significant. Koul et al. [202] highlighted the energy-intensive nature of conventional wastewater treatment and the potential of MECs for resource recovery. They advocated for integrating MECs with anaerobic digestion and membrane bioreactors to enhance energy efficiency and sustainability. Radhika et al. [203] reviewed various MEC applications, including energy generation, methane and hydrogen peroxide production, and pollutant removal. Their study also discussed advancements in biocathode materials, reactor design, and coupling MECs with microbial fuel cells for synergistic effects.
Cho et al. explored the use of the waste electrolysis cell (WEC) to decentralize hydrogen production with onsite water treatment. The WEC consisted of the stainless steel cathode and the multi-junction semiconductor anode in a single compartment cell. The highest energy efficiency (EE = 0.23) and current efficiency (CE = 0.8) for HER were observed in electrolysis at current densities of 200 A/m2 of real wastewater [204]. Lu et al. produced a NiFe layered double hydroxide (NiFe LDH) electrocatalyst on NF for H2 production from wastewater. The new cathode showed a high H2 rate with the Pt catalyst, which was twice as high compared to stainless steel and bare NF cathodes [205]. Kaider et al. examined a metal electroformed Ni mesh cathode alternative to Pt/CC in a single chamber membrane-free MEC. According to results, it was concluded that great potential should be used in Ni mesh catalyst as a cathode material for H2 production in a single chamber membrane-free MEC [206]. Lim et al. stated the bioanode at a current density of 0.36 A/m2 and 0.37 A/m2 with the potential of 0 and +0.6 V, respectively. During this, the H2 production was 7.4 L/day at a cathode potential of −1.0 V [207]. Chen et al. improved the biocathodes with PANI (polyaniline)/MWCNT composites for higher H2 production in single chamber, membrane-free biocathode MECs. The results concluded that with an increment in applied voltage, the H2 production rates increased, and the performance of MECs was improved with modified biocathodes at a higher current density and H2 generation rate [208]. Li et al. obtained excellent perchlorate reduction under various initial concentrations in a non-membrane MEC with a PANI-modified graphite cathode as the sole electron donor [209]. Liu et al. used microbial fluidized electrode electrolysis cells to increase the H2 gas production. In this, the flowable granular activated carbon (GAC) particles were used for the growth of exo-electrogenic bacteria, which enhanced hydrogen production in bio-electrochemical systems [210]. Wang et al. evaluated the feasibility of operating the MEC at low temperatures (10 °C) using biocathodes. It was stated that H2 could be generated from wastewater by using biocathodes [211]. Huang et al. observed that a self-driven microbial electrolysis cell or microbial electrolysis cell can completely release Co (II) from LiCoO2 on the cathodes in MFCs. This study provided a new process of linking MFCs to MECs to recover cobalt and recycle it to Li-ion batteries with no external energy consumption [212]. Cusick et al. designed a two-chamber MEC to produce suspended particles and discovered that MEC is an expectant method of energy recovery and electrochemical nutrient for nutrient-rich wastewaters [213]. Figure 13a shows the water electrolysis cell with the plot showing current efficiency and electrical efficiency [204]. Figure 13a demonstrates redox and mass transport processes involving water splitting, chloride oxidation to free chlorine, and degradation of COD, enhancing both hydrogen production and pollutant removal. Efficiency graphs indicate that while current efficiency increases with current density, energy efficiency remains relatively low. Material analysis confirms the electrode composition (Bi, Ti, O) and favorable morphology for electrochemical activity. Figure 13b shows the plot between potential and time at different cell operating conditions [212]. Figure 13b includes the lab setup and a performance graph showing electrode potentials over five days under different cathodic biases. A sharp change around day 3.5 indicates bioanodic performance loss in MEC, emphasizing the importance of biofilm stability for long-term operation. Marone et al. evaluated bio-H2 production by dark fermentation and microbial electrolysis from agro-industrial wastewater. The results concluded that the above combination was a promising option for optimizing the conversion of wastewater [214]. Huang et al. have used toilet wastewater for MEC. Their primary goal was to establish the feasibility of MEC for toilet wastewater disinfection [215]. Dhar et al. produced hydrogen from sugar beet juice (SBJ) using MEC. The H2 production was 25% of the initial COD, and the energy recovery from SBJ was 57% by combined biohydrogen [216]. Heidrich et al. performed MEC using domestic wastewater. The 100-L MEC was operated on raw domestic wastewater at 1 °C to 22 °C, producing 0.6 L/day of H2 [217]. Trably et al. used saline wastewater in a biofilm-based 4 L two-chamber MEC continuously fed with acetate under saline conditions for more than 100 days [218]. Li et al. investigated the integrated microbial desalination cell-MEC system by saving nitrogen, metal, and saline from municipal wastewater, industrial wastewater, and seawater, respectively, to observe the nitrogen degradation rate [219]. Kuntke et al. have performed MEC using urine for ammonium removal and hydrogen production [220].
Linji et al. mixed trehalose with wastewater to perform MEC. The feasibility of physical and biological processes was observed for sludge treatment and the effects of trehalose on the H2 generation of MEC at 0 °C [221]. Sun et al. recovered hydrogen from high solid waste activated sludge (WAS) using MEC. With the optimal concentration, maximum hydrogen yields were reached for MECs fed with raw WAS and alkaline preheated WAS [222]. Zhang et al. investigated the improvement of a zero-valent iron-activated carbon (ZVI-AC) micro-electrolysis system on H2 production by a mixed bacterial consortium. The results showed that 38.2% more hydrogen was produced with the addition of ZVI, and further addition of ZVI-AC enhanced the H2 production [223]. Wu et al. performed electrolysis of 0.33 M urea using a Ni hydroxide electrode on stainless steel foil by cathodic electrodeposition with 1 M KOH electrolyte. It was observed that the Ni hydroxide electrode obtained much better electro-catalytic performance than the film and hollow sphere electrodes [224]. Figure 14a shows a two-stage process combining dark fermentation with an MEC, enhancing hydrogen yield by converting fermentation products into H2 electrochemically [214]. Figure 14b integrates MECs with solar power, showcasing a renewable-driven setup with improved energy and environmental performance [215]. Figure 14c features sugar beet waste utilization in a hybrid system yielding high H2 efficiency and energy recovery [216]. Figure 14d highlights the seasonal dependence of MEC performance in outdoor conditions, with higher efficiency during warmer months due to elevated microbial activity [217]. Figure 14e demonstrates a continuous-flow MEC using granular bioanodes and membrane separation to achieve 90% H2 purity, emphasizing real-world viability [218]. Finally, Figure 14f illustrates the competition between hydrogen and methane production, underlining the importance of microbial control for selective H2 generation [219]. MEC performance and stability depend upon the experimental setup. Zhang et al. have used a double anode arrangement to perform MEC for achieving H2 production from glucose by adding CH4. It was observed that the chloroform was an effective methane inhibitor and improved the efficiency of H2 production from glucose in the MECs [225]. Lewis et al. used integrated pyrolysis- MEC to describe hydrogen production. The results demonstrated that the pyrolysis microbial electrolysis process was a sustainable and efficient route for the production of H2 from biomass [226]. Watson et al. used a microbial reverse-electrodialysis electrolysis cell (MREC) from the fermentation of wastewater to yield H2. The results confirmed that if effluent anolyte COD concentrations were sufficient for anode potential, then the consistent rates of H2 produced by MREC [227]. Song et al. have used substrate (without buffer solution) under continuous flow conditions in an MREC for H2 production. The hydrogen produced was 0.61 m3-H2, with a COD removal efficiency of 81% and a coulombic efficiency of 41% in MREC [228].
Chen et al. investigated the thermoelectric micro-converter MEC coupled system for H2 production from acetate. This study showed that this system directly converted waste heat energy to electricity at a temperature difference of 5 °C and helped MEC to produce H2 [229]. As a monitoring tool for hydrogen-producing MEC, Montpart et al. presented a low-cost sensor for H2 production measurement. As a result, the fuel cell electrical signal had a high correlation with H2 production, and fuel cell quantification was proved to be equivalent to gas chromatography analysis [230]. Guo et al. reported on a liter-scale tubular MEC for obtaining high rate H2 production; its components, such as a Pt-coated Ti mesh cathode, an anion exchange membrane, and a pleated stainless steel felt anode, were arranged in a concentric configuration. The reactor presented high H2 recovery, high H2 purity, and outstanding operational stability [231]. Montpart et al. have also used a single chamber MEC for H2 production from synthetic wastewater. The results displayed that at 0.8 V, the current intensity, H2 production, and cathodic gas recovery were 150 A/m2, 0.94 H2m3/m3d, and 91%, respectively [232]. Lou et al. utilized a dual-chamber MEC for H2 production, and also metal removal, such as Cu2+, Ni2+, and Fe2+ from acid mine drainage (AMD) [233]. Zhang et al. described an active iron-reducing bacteria (IRB) through dosing Fe (III) into an MEC for the degradation of organic matter. This study provided a method to enrich electrochemically active IRB in the bio-electrochemical reactor for treating industrial wastewater [234]. The membrane affects H2 production in electrolysis, so even in MEC, the effect of the membrane on electrolysis was reported. Chae et al. prepared a nanofibre-reinforced composite proton exchange membrane based upon a proton conductor, i.e., sulfonated polyether ketone (SPEEK). This novel membrane was compared with the Nafion membrane and resulted in lower gas and fuel crossovers with higher proton conductivity that improved H2 production at the cathode, with overall H2 efficiency as compared to Nafion [235].
Khan et al. reviewed the status of MEC as a mean wastewater treatment method and H2 production. This study estimated a total electricity of 434 MWe could be produced in 2015 from the Kingdom of Saudi Arabia’s wastewater if MEC technology was employed [236]. Cotteril et al. reviewed the production of H2 in MEC and treated the wastewater to reduce the energetic and economic costs of operation [237]. Zhang et al. further discussed the recent advances and future challenges of MEC. This approach significantly reduced the cost of the electric energy for H2 production as compared to direct water electrolysis [238]. Kaider et al. have given a review of the substrates used in MEC, as there is a large number of substrates that could be used as a fuel source [239]. Rago et al. analyzed the microbial community in membrane-less MEC. As a single chamber MEC, the production of H2 failed due to methanogenesis buildup. Moreover, at higher 2-BromoEthanesulfonate concentrations, the methanogenesis activity was decreased and increased the homo acetogenesis activity, which optimized the performance of the MEC for H2 production [240]. Kaider et al. completed a review of reactor design and configuration of MEC. The study showed that MEC reactor design directly influences the H2 and current production rate in MECs, because the membrane-free design can lead to both on H2 production rate and recovery rates [241]. Further, the recent advances and challenges in MEC were also reviewed [242].
Significant advancements have been made in MEC research, particularly in the development of alternative electrode materials to replace expensive platinum-based catalysts (Table 6). Materials such as NiFe layered double hydroxides (LDH), polyaniline (PANI)/MWCNT composites, and Ni mesh electrodes have demonstrated enhanced hydrogen production. System configurations have also evolved, with studies exploring single chamber, membrane-free designs and microbial reverse-electrodialysis electrolysis cells (MREC) to improve efficiency and reduce operational costs. MECs have been integrated with other technologies to optimize hydrogen production and wastewater treatment. For example, microbial desalination cells, thermoelectric converters, and dark fermentation processes have been combined with MECs to enhance overall performance. Additionally, research on optimizing operating conditions, such as temperature, pH, applied voltage, and microbial communities, has provided insights into improving system efficiency. Studies on low temperature MECs have demonstrated feasible hydrogen production at 10 °C, making the technology applicable in colder regions.
Despite these promising advancements, MECs still face critical challenges related to high internal resistance, low hydrogen purity, interference from methanogenesis, and scalability constraints. Overcoming these limitations necessitates improvements in biocatalyst efficiency, reactor design optimization, and comprehensive techno-economic assessments. While research has advanced from proof-of-concept studies to refined designs featuring enhanced electrode materials and system efficiencies, commercial deployment remains hindered by both technical and economic barriers.
Recent studies have demonstrated that alternative electrode materials, such as NiFe LDH and Ni mesh cathodes, can reduce costs while maintaining efficiency, while single chamber and membrane-free configurations provide advantages in hydrogen recovery and system simplicity. Process integration with dark fermentation and thermoelectric conversion has further enhanced MEC performance, improving overall energy yields. However, achieving scalability, cost-effectiveness, and long-term operational stability remains a major challenge. Future advancements must focus on developing durable, cost-effective, and high efficiency catalysts to replace expensive noble metals. Hybrid systems, integrating MECs with other bio-electrochemical and renewable energy technologies, such as photovoltaic-MEC hybrids, could improve overall energy recovery and system viability. Additionally, automation and AI-driven real-time monitoring could optimize MEC performance, energy management, and maintenance strategies.
Scaling up MECs for industrial and municipal wastewater treatment is another critical research direction, as these applications could leverage the dual benefits of sustainable hydrogen production and organic waste remediation. However, comprehensive techno-economic analyses are needed to determine the feasibility of large-scale implementation. By addressing these scientific and engineering challenges, MECs have the potential to transition from laboratory research to commercially viable solutions for hydrogen production and wastewater treatment, contributing significantly to the water-energy-resource nexus.

4. Advancements and Emerging Trends in Electrolysis Technologies for Hydrogen Production

Hydrogen production via electrolysis is influenced by secondary factors beyond electrode and electrolyte selection, including gas bubble dynamics, mass transport limitations, system efficiency losses, and integration with renewable energy sources. Addressing these challenges is essential for scaling up electrolysis technologies and improving long-term performance. Recent advancements focus on controlling bubble formation, optimizing electrochemical interfaces, enhancing catalytic efficiency, and refining membrane extraction methods. Additionally, alternative electrolysis approaches, such as high temperature, ionic-liquid-assisted, and hybrid electrolysis, are being explored to improve hydrogen yield, energy efficiency, and cost-effectiveness. These developments pave the way for more efficient and commercially viable electrolysis systems. Chandran et al. experimentally investigated the H2 bubble formation characteristics and its transport near the electrode surface. It was concluded that the bubble velocity was key in defining the bubble size distribution near the electrode surface [243]. Wang et al. further developed a method to determine the effect of the gas bubble on cell voltage oscillations. The simulation on Matlab/Simulink revealed that the gas bubbles on the cell voltage exceeded the normal fluctuation amplitude [244]. Goupil et al. conducted a comprehensive study on the behavior of CuNiFe anodes and CuNiFeO electrodes at 700 °C in potassium cryolite-based electrolyte during the electrolysis [245]. Gutiérrez-Martín et al. explored hydrogen and power integration strategies through large-scale water electrolysis in the Spanish power system, highlighting its potential for flexible energy storage and grid balancing. However, the study reveals several challenges, including the high sensitivity of hydrogen production costs to electricity prices, underutilization of electrolyzers due to fluctuating renewable surplus, and complex grid integration requirements. The economic feasibility is further constrained by high capital costs and the inefficiencies of part-load operation. The authors suggested strategies such as dynamic operation aligned with real-time electricity pricing, decentralized electrolyzer deployment, co-location with renewable sources, and integration with fuel cells for improved energy return [246]. Alhahosseini et al. investigated the challenges associated with the Cu-Cl cycle of H2 production. In the Cu–Cl cycle, hydrolysis of CuCl2 requires a significant excess of steam (up to 17:1 H2O to CuCl2 molar ratio) to ensure high yields of Cu2OCl2, but this results in a diluted HCl gas product, which is insufficient for the subsequent electrolysis step that demands 6–11 M HCl for optimal efficiency. To address this, the authors propose an integrated system involving a heat recovery steam generator (HRSG) and an HCl–H2O separation process using rectification and absorption columns. Their simulation demonstrates that this integration not only supplies the required concentrated HCl (up to 22 mol%) for electrolysis but also recovers excess steam to improve the thermal efficiency of the cycle. The study highlights that with effective integration, the Cu–Cl cycle can achieve thermal efficiencies as high as 88.6%, making it a competitive alternative to conventional water electrolysis [247]. Liu et al. investigated coal pretreatment using the ionic liquid [Bmim]Cl. The treatment disrupted coal’s structure, increasing porosity and breaking hydrogen bonds. This structural modification enhanced access to reactive sites during electrolysis. As a result, electro-reduction activity improved by 15% compared to untreated coal [248]. Chen et al. have investigated the effects of ionic liquids, which included the concentration, the structure, the temperature, and the time, on the removal of sulfur from coal water slurry electrolysis. The results showed the effect of the anions on desulfurization and exhibited better performance as compared to pyridine [249]. Slampova et al. studied that electrolysis can be controlled by proper electro-membrane extraction method (EME) selection for H2 production [250]. Cheng et al. summarized the progress of metal oxides catalysts and the role of carbon nanotubes (CNTs) in OER catalysts with the latest development [251]. Guerra et al. described the production of synthesis gas by using graphite electrodes from water electrolysis without the separation of the produced gases. The obtained synthetic gas can be used for fuels, such as CH4, DME (dimethyl ether), or methanol in a catalytic reactor, using the ELECTROFUEL concept [252].
Anion exchange membrane water electrolyzers (AEMWEs) have emerged as a promising alternative to conventional electrolysis technologies, aiming to combine the cost advantages of alkaline water electrolysis with the efficiency of PEM systems. Operating in an alkaline environment while employing a solid polymer electrolyte, AEMWEs enable the use of non-precious metal catalysts, maintain high ionic conductivity, and support compact system designs. As Peng [253] and Sugawara et al. [254] note, their performance is critically governed by the membrane electrode assembly (MEA), where synergy between the membrane, ionomer, and electrocatalysts determines overall efficiency.
Recent research has addressed several key challenges in AEMWE development. Li et al. [255] enhanced OER kinetics and long-term stability using chromium-doped amorphous oxides (e.g., CoCrOx), while Lim et al. [256] introduced a quaternized polyfluorene-based membrane (PFPBPF-4-QA) with high conductivity and alkaline stability. Zheng et al. [257] demonstrated a system-level breakthrough by achieving stable operation at ultrahigh current densities (10 A/cm2 for over 800 h) through integrated optimization of membranes, ionomers, and porous transport layers.
Catalyst development has also advanced. Klingenhof et al. [258], using operando spectroscopy, explored Ni–O redox-active ligand systems, revealing mechanisms that enable PEM-like OER performance without noble metals. Xu et al. [259] and Chang et al. [260] contributed insights into interface engineering and numerical modeling, offering strategies to improve mass/charge transfer and apply materials innovations to real-world system configurations.
On the durability front, Motealleh et al. [261] reported over 10,000 h of stable operation using Sustainion® membranes, achieving degradation rates below 1 mV/h—a notable benchmark for commercial viability. Liu et al. [262] complemented these findings by identifying chemical degradation pathways and proposing structural strategies to improve long-term conductivity and alkaline resistance.
Collectively, these studies demonstrate the rapid progress of AEMWEs through innovations in membrane design, catalyst engineering, and system integration. Their inclusion in this review strengthens its relevance by capturing one of the most promising directions in next-generation water electrolysis.
Advancements in water electrolysis technologies have largely been driven by material innovations, process optimizations, and experimental studies. While experimental research provides critical insights into electrode materials, electrolyte composition, and operating conditions, it often faces challenges in capturing microscopic charge transport phenomena, reaction kinetics, and mass transport limitations. This limitation makes it essential to complement experimental findings with computational modeling techniques, which allow for a deeper understanding of electrochemical interactions, system efficiencies, and degradation mechanisms.

5. Numerical Modeling and Finite Element Analysis of Water Electrolysis

Computational modeling techniques such as the finite element method (FEM), computational fluid dynamics (CFD), response surface methodology (RSM), and system-level models (e.g., Aspen Plus®) play a critical role in optimizing water electrolysis systems. These methods allow for the analysis of electrochemical kinetics, thermal distribution, fluid dynamics, and performance degradation. Each modeling strategy offers unique strengths depending on the scale and objectives of the analysis, as summarized in Figure 15.
As shown in Figure 15, FEM is commonly used for localized, detailed simulations involving mechanical stress and temperature distribution. CFD, on the other hand, models multi-phase flow and bubble dynamics in electrolyzer cells. RSM provides empirical optimization of multiple variables with fewer simulation runs, while Aspen Plus® supports techno-economic and system-level analyses for full-scale integration of electrolyzers with energy sources.
Each numerical simulation method used in water electrolysis research serves a distinct modeling purpose, and its applicability depends on the scale, objective, and complexity of the problem being addressed. The FEM is best suited for detailed, cell-level investigations where coupled physical phenomena such as electrochemical kinetics, thermal gradients, and mechanical stress coexist. Its ability to solve coupled partial differential equations over complex geometries makes it ideal for analyzing localized effects in catalyst layers, membrane–electrode assemblies, and current collectors. FEM is particularly effective for optimizing component configurations and predicting degradation behavior, albeit at a high computational cost.
In contrast, CFD is the preferred tool when analyzing mass transport, bubble dynamics, or flow field design. CFD excels in simulating two-phase flow phenomena in alkaline and PEM electrolyzers, especially under high gas evolution rates. While it lacks the intrinsic electrochemical modeling fidelity of FEM, CFD provides essential insight into gas-liquid interactions, which influence ohmic resistance and bubble-induced overpotentials. For broader design-of-experiment applications, RSM offers an efficient statistical framework for optimizing performance across multiple operating parameters. Though it does not resolve spatial or temporal physics, RSM is valuable for identifying influential variables and constructing surrogate models for process optimization.
System-level models, such as those developed in Aspen Plus®, are indispensable for analyzing integrated hydrogen production systems. These models handle energy and mass balances, load-following dynamics, and techno-economic parameters, making them highly relevant for coupling electrolyzers with renewable sources such as solar or wind. However, Aspen Plus® operates at the plant scale and does not capture internal electrochemical or transport dynamics. Finally, multiphysics coupling platforms (e.g., COMSOL Multiphysics) provide a unified environment to simulate thermal, fluidic, and electrochemical effects concurrently. These tools are ideal for capturing cross-domain phenomena in advanced electrolyzer designs but demand significant computational resources and model development expertise.
In summary, FEM offers the most comprehensive physical fidelity for cell-level simulations, CFD dominates in hydrodynamic analysis, RSM simplifies parameter optimization, Aspen Plus® enables system-scale integration, and multiphysics frameworks offer synergistic modeling of interacting phenomena. Selecting the appropriate method requires careful consideration of modeling depth, computational efficiency, and the specific challenges addressed in the electrolyzer system under study.
While various modeling methods have been explored, FEM is emphasized in greater detail due to its wide applicability in stress analysis, thermal management, and multiphysics coupling within electrolyzer components. Numerical modeling, particularly FEM-based simulations, has emerged as a powerful tool to bridge the gap between experimental observations and theoretical predictions by providing quantitative and predictive analyses of electrochemical behavior. Given the multi-scale, coupled transport phenomena and complex electrochemical interactions in PEM electrolysis, it was selected as the focus of this numerical study. Unlike AWE, which operates with well-understood liquid-phase mass transport, PEM electrolysis involves interdependent kinetic, thermodynamic, and mass transport processes that significantly influence system performance. These characteristics make PEM an ideal candidate for FEM-based electrochemical simulations, enabling precise performance predictions under varying operational conditions.
Numerical modeling and simulations play a vital role in evaluating design changes, predicting performance, and avoiding costly experimental iterations. FEM and CFD are increasingly employed to optimize the design and performance of AWE systems by enabling detailed analysis of fluid flow, gas evolution kinetics, current density distribution, heat transfer, and structural mechanics. The fidelity of FEM and CFD simulations in electrolysis systems depends critically on the accuracy of input parameters such as electrolyte conductivity, electrode geometry, membrane transport coefficients, gas-liquid interface behavior, and heat transfer properties. When properly calibrated against experimental data, high-fidelity models can predict key performance metrics (e.g., current density, temperature gradients, pressure drop) within 5–10% of measured values. However, challenges remain in simulating bubble-induced resistance, local pH shifts, and two-phase flow transients. To enhance model robustness, advanced simulation workflows incorporate mesh refinement studies, sensitivity analyses, and partial experimental validation to improve predictive confidence and ensure reliability in design-oriented applications.
Several studies underscore the importance of integrating optimized geometrical configurations, mechanical assembly parameters, and operational conditions to maximize the performance and durability of PEM and alkaline water electrolyzers. Yang et al. [263] investigated a novel orthogonally sinusoidal flow channel geometry for PEMECs, finding that increased amplitude and angular frequency in the horizontal and vertical planes significantly improved electrochemical performance and mass transport. Similarly, Tirumalasetti et al. [264] compared multiple flow field designs and demonstrated that the Double-Layered Wire Mesh (DLWM) configuration enhanced hydrogen mole fraction and current density distribution more effectively than conventional flow fields. The optimization of structural and operational parameters plays a crucial role in the performance and longevity of electrolyzers. Kink et al. [265] conducted finite element simulations to evaluate mechanical stresses on PEM membranes and found that membrane buckling could be prevented by maintaining a gap below 0.3 mm between the membrane and the porous transport layer. Complementarily, Ozdemir et al. [266] applied response surface methodology alongside finite element modeling to optimize bolt torque in PEMWE assemblies. They identified an optimal torque of 10 Nm, which minimized mechanical stress while preserving membrane integrity and electrochemical efficiency. Operational conditions such as pressure, temperature, and flow rate also significantly influence system performance. Hassan et al. [267] used a 3D non-isothermal model to analyze the effects of gas diffusion layer thickness and porosity, reporting that higher temperatures improved performance while elevated cathode pressures impaired hydrogen ion diffusion and decreased efficiency. Jang et al. [268] explored pressure effects in an alkaline water electrolysis system using Aspen Plus® modeling and showed that high pressure operation improved hydrogen purity and reduced energy consumption, although purity gains diminished above 20 bar. Transport phenomena and flow field design continue to be pivotal in enhancing performance. Zheng et al. [269] used 3D simulations to analyze various flow field patterns and concluded that parallel flow channels provided superior current distribution and minimal pressure drop. Their work highlighted the importance of temperature and flow structure in facilitating efficient mass transfer.
To further enhance the understanding of PEM electrolysis, this study employs an FEM-based approach to model key electrochemical and mass transport processes, allowing for precise prediction of system performance, efficiency, and degradation mechanisms. Specifically, we analyze electrochemical potential distribution, overpotential contributions, mass transport dynamics, and gas bubble evolution effects. By incorporating a multi-physics simulation framework, the study provides valuable insights into reaction kinetics, ion transport, and electrode behavior, offering guidance for performance optimization and system design improvements. This numerical modeling approach not only enhances the fundamental understanding of PEM electrolysis but also serves as a predictive tool for optimizing key parameters such as current density, electrode material selection, and pressure settings, ensuring higher hydrogen production efficiency and system longevity.

5.1. Finite Element Model for PEM Electrolysis

The FEM-based approach employs numerical solutions to Laplace’s equation for electrochemical potential distribution, Nernst–Planck equations for ion transport, and Butler–Volmer kinetics for electrode reactions. The key governing equations include Electrochemical Potential Distribution (Laplace’s equation) that governs charge transport across the electrolyte and electrodes. The electrochemical potential distribution follows Equation (1).
2 = 0
where:
  • ∅ = electrochemical potential (V).
Mass Transport (Nernst–Planck Equation): Models ion diffusion and electromigration in the electrolyte:
The transport of ions is modeled as Equation (2).
· ( D c + μ c ) = R
where:
  • C = ion concentration (mol/m3)
  • D = diffusion coefficient (m2/s)
  • μ = ion mobility (m2/V·s)
  • R = reaction rate (mol/m3·s)
c t = D 2 c · ( z F c E c )
where:
  • E = the electric field.
Electrode Reaction Kinetics (Butler–Volmer Equation):
j = j 0 e α a F η / R T e α c F η / R T
where:
  • j = current density
  • j0 = exchange current density
  • η = activation overpotential.
These equations were solved using FEniCSx 0.8. (open-source FEM solver) with a 3D cubic mesh (32 × 32 × 32 elements), providing insights into spatial variations in electrochemical properties.

5.2. Gas Evolution and Bubble Dynamics

One of the major challenges in PEM electrolysis is bubble formation at electrode surfaces, which introduces resistance to ion transport and reduces reaction efficiency. The FEM model integrates a multiphase CFD approach, solving the Navier-Stokes equations coupled with a Volume of Fluid (VOF) model to track bubble growth and detachment. The effective electrolyte conductivity (σeff) is modeled as:
σ e f f = σ b u l k ( 1 ϕ )
where:
  • ϕ = gas volume fraction, reducing ion transport efficiency as bubbles accumulate.
Gas Bubbles Introduce Resistance to Ion Transport:
σ e f f = σ ( 1 ϵ )
where:
  • σ = bulk electrolyte conductivity
  • ϵ = gas volume fraction.
ϵ = k b j b
In Equation (7), j is the local current density (A/m2), and kb and b are empirical fitting constants that depend on electrolyte type and operating conditions. These are used to estimate the gas volume fraction ϵ, which quantifies average gas holdup and its impact on effective conductivity.
Modified Ohm’s Law with Gas Bubble Effects:
J = σ e f f
Electrode Reaction Kinetics (Butler–Volmer Equation)
The reaction kinetics are modeled using:
j = j 0 e α a n F η / R T e α c n F η / R T
η = r e v i R
In Equation (9), j is the local current density (A/m2), j0 is the exchange current density, αa and αc are the anodic and cathodic charge transfer coefficients, F is the Faraday constant (96,485 C/mol), η is the overpotential, R is the universal gas constant (8.314 J/mol·K), and T is the temperature in Kelvin. Equation (10) defines the overpotential η, where ϕ is the local electric potential, ϕrev is the reversible potential, and iR accounts for ohmic losses.
While both ϕ and ϵ are referred to as gas volume fractions, they are used in distinct modeling contexts: ϕ represents the local, spatially resolved gas fraction typically used in FEA/CFD-based simulation, whereas ϵ is a bulk or averaged gas holdup used in empirical conductivity correction models.

5.3. 3D Finite Element Method Implementation

To enhance the accuracy of the electrochemical and mass transport simulation, a 3D FE model is implemented using the FEniCS open-source FEM solver. The inclusion of 3D modeling allows the study of spatial variations in electrochemical potential, current density distribution, and reaction kinetics that are not captured in a simplified 2D approach.

5.3.1. Computational Domain and Discretization

A cubic computational domain is used, representing a unit segment of the PEM electrolyzer. The domain is meshed into 32 × 32 × 32 finite elements, ensuring sufficient resolution for electrochemical and mass transport processes. The mesh is refined near the electrode-electrolyte interface to improve accuracy in regions of high reaction activity. This resolution was selected based on a mesh independence test to ensure numerical stability and computational efficiency. The model represents a simplified unit domain of a PEM water electrolysis cell under steady-state conditions, assuming uniform material properties and isotropic conductivity. The mesh is sufficiently fine to resolve key electrochemical gradients—particularly near the catalyst–membrane interface—without introducing significant error in current density or potential profiles. Regions of high reaction intensity were verified to exhibit smooth potential distribution, supporting the mesh adequacy for the intended resolution of local ohmic and activation losses.

5.3.2. Boundary Conditions in 3D Simulation

  • Anode (x = 0): Fixed potential of 1.23 V (standard water splitting potential)
  • Cathode (x = L): Grounded at 0 V
  • Other Boundaries (y and z directions): Insulated, with zero normal flux (Neumann conditions)
  • Electrode Reaction Interface: Butler–Volmer boundary conditions applied at electrode surfaces to model the charge transfer reaction kinetics.
Although the anode (x = 0) and cathode (x = L) were set to 1.23 V and 0 V, respectively, for solving Laplace’s equation, the model does not neglect cell overpotential. Instead, the activation, ohmic, and concentration overpotentials are separately computed and superimposed to simulate realistic cell performance. These include contributions from charge transfer kinetics (via the Butler–Volmer equation), electrolyte conductivity losses, and mass transport limitations—ensuring the overall electrochemical losses are comprehensively captured.

5.3.3. Numerical Solution Strategy

The weak form of the governing equations is derived and solved using Galerkin FEM discretization. The nonlinear system is solved iteratively using the Newton-Raphson method. The direct linear solver (UMFPACK) is employed for high-precision numerical solutions. A grid independence study is conducted to ensure numerical accuracy. The assumptions and parameters for the PEM electrolyzer are given in Table 7.
The values used in Table 7 represent commonly reported baseline parameters for PEM electrolyzer modeling under standard operating conditions (e.g., 60 °C, 1 atm, acidic media). These values were selected based on consistency with multiple simulation studies and general consensus in the literature. Where necessary, parameters were adjusted to match experimental calibration and ensure numerical stability in the FEM implementation.

5.3.4. Results from 3D FEM Implementation

The 3D model captures non-uniformity in electrochemical potential across the domain, providing a realistic assessment of performance variations within the PEM electrolyzer. Higher current density is observed near the electrode edges, aligning with experimental studies. The impact of bubble formation on localized resistance and mass transport limitations is better visualized in 3D. This detailed 3D FEM implementation significantly enhances model fidelity, allowing a more accurate representation of PEM electrolysis under real-world conditions.
The electrochemical potential distribution within the electrolyte, as shown in Figure 16, provides critical insight into the charge transport behavior across the PEM electrolyzer. This figure is obtained by solving Laplace’s equation for electrochemical potential, considering the applied boundary conditions at the electrodes. The anode is set at 1.23 V, corresponding to the standard water splitting potential, while the cathode is grounded at 0 V. The numerical results illustrate a smooth voltage gradient across the electrolyte domain, confirming uniform charge transport under steady-state conditions.
In the 2D model, the potential distribution follows a predictable linear decline, with minor variations due to localized resistance variations at the electrode-electrolyte interface. The 3D model, however, reveals more complex spatial variations, particularly at the electrode edges, where non-uniform current density distributions arise. This behavior is attributed to the geometric effects of the electrode structure and the influence of localized reaction kinetics, which are more accurately captured in the 3D FEM model. The voltage gradient confirms that ohmic losses within the electrolyte are the primary contributors to internal resistance, a conclusion that aligns well with experimental studies. These results validate the numerical approach’s accuracy in simulating realistic charge transport behavior, making it a reliable tool for optimizing electrode configurations and electrolyte properties in practical PEM electrolyzers.
The breakdown of overpotential contributions, depicted in Figure 17, illustrates the role of activation, ohmic, and concentration overpotentials in governing overall electrolyzer efficiency. This figure is generated by computing individual overpotential contributions based on the Butler–Volmer equation (activation), Ohm’s law (ohmic), and the Nernst–Planck equation (concentration effects). The numerical results closely align with experimental polarization data, reinforcing the accuracy of the model.
The activation overpotential, which dominates at low current densities, arises due to the sluggish OER kinetics at the anode. This overpotential follows an exponential dependence on current density, as dictated by the charge transfer kinetics in the Butler–Volmer equation. As the current density increases, the ohmic overpotential becomes the dominant factor, exhibiting a linear dependence on current density, in agreement with Ohm’s law. This behavior is well captured by the model, with deviations of less than 2% from experimental results [270,271], confirming the accuracy of the electrolyte conductivity and cell resistance assumptions. At very high current densities, the concentration overpotential becomes noticeable, attributed to mass transport limitations in the electrolyte. This effect is modelled using the Nernst–Planck transport equation, capturing the impact of ion depletion near electrode surfaces. The slight deviations in concentration overpotential (~5%) compared to experimental results [270,271] may be due to bubble formation effects, which introduce additional transport resistance. Overall, the agreement between the numerical and experimental overpotential trends confirms that the model provides an accurate representation of electrochemical losses in PEM water electrolysis.
The final gas fraction distribution, illustrated in Figure 18, highlights the impact of bubble formation on mass transport and electrolyte conductivity. The gas fraction distribution is obtained using a multiphase CFD approach, which tracks the growth and detachment of hydrogen and oxygen bubbles within the electrolyte. These bubbles form as a result of the electrochemical splitting of water, and their accumulation affects local conductivity and ion transport dynamics. The simulation considers gas-phase transport equations, incorporating bubble nucleation rates, growth mechanisms, and detachment criteria to predict the spatial distribution of gas fractions. The results indicate that gas fraction is highest near the electrode surfaces, where gas evolution occurs. As bubbles grow and detach, they introduce localized regions of reduced conductivity, increasing electrolyte resistance and affecting overall cell performance. The model captures the spatial variation of bubble-induced transport resistance, showing that regions with higher gas fraction exhibit higher ohmic losses.
In 3D simulations (Figure 19), gas accumulation effects are more pronounced, particularly near the electrode edges and flow channels, where bubble detachment rates vary due to local hydrodynamic conditions. The final gas fraction distribution aligns well with experimental observations, showing that bubble-induced resistance must be accounted for in practical PEM electrolyzer designs. The results suggest that strategic electrolyte flow control and electrode surface modifications can mitigate the negative impact of gas accumulation, optimizing system efficiency. The model provides an accurate predictive tool for evaluating gas evolution effects in large-scale electrolyzer systems, demonstrating its potential for performance enhancement and design optimization.
The comprehensive analysis of electrochemical potential distribution, overpotential contributions, and gas fraction effects provides crucial insights into the internal mechanisms governing PEM water electrolysis. The observed trends in voltage gradients, activation losses, and bubble-induced resistance directly influence overall cell performance and dictate key operational parameters. Building on these findings, the discussion now extends to the polarization behavior of the electrolyzer, which serves as a macroscopic performance indicator by relating cell voltage to current density. Furthermore, the hydrogen production rate—a fundamental measure of electrolyzer efficiency—is examined in the context of Faraday’s Law and reaction kinetics. Finally, to assess system performance holistically, the efficiency vs. current density relationship is explored, identifying operational regimes where the electrolyzer achieves optimal energy utilization. Together, these evaluations provide a detailed framework for understanding both the internal electrochemical dynamics and the practical performance characteristics of the system.
The polarization curve represents the fundamental relationship between current density and cell voltage, serving as a critical performance indicator for PEM water electrolysis. This curve is derived by solving the coupled electrochemical potential equation (Laplace’s equation) and reaction kinetics (Butler–Volmer equation) under varying current densities. The numerical results reveal a characteristic increase in cell voltage with increasing current density, attributed to three primary overpotential contributions: activation, ohmic, and concentration overpotentials. At low current densities, the activation overpotential dominates due to the inherent sluggishness of the OER at the anode. This effect is well captured by the Butler–Volmer equation, which accounts for the exponential dependence of the charge transfer kinetics on the applied overpotential. As the current density increases, the ohmic overpotential becomes the major contributor to voltage rise, governed by Ohm’s law and dependent on the electrolyte conductivity and electrode resistance. The model accurately captures the linear increase of ohmic losses, validating its agreement with experimental data. At very high current densities, concentration overpotentials become noticeable, primarily due to mass transport limitations in the electrolyte. This results from the depletion of reactant ion concentrations near the electrode surfaces, which the model incorporates by solving the Nernst–Planck equation for ion transport.
The polarization curve (Figure 20a) illustrates the relationship between cell voltage and current density, serving as a critical indicator of electrolyzer performance. The breakdown of overpotential contributions across different current densities further validates the model’s accuracy. Additionally, overpotentials contribute to the difference between the thermodynamic and actual cell voltages. The activation overpotential follows an exponential trend (Figure 20b), while the ohmic overpotential scales linearly, and the concentration overpotential remains minimal but rises at higher current densities. A comparison with experimental studies shows that the activation overpotential deviates by ~5%, likely due to approximations in exchange current density and reaction kinetics. Meanwhile, the ohmic overpotential matches experimental values within <2% deviation, demonstrating the reliability of the conductivity and resistance assumptions.
The hydrogen production rate is another critical performance parameter, derived using Faraday’s Law, which establishes a direct relationship between current density and hydrogen generation rate. The model accurately predicts that hydrogen production increases linearly with current density, consistent with electrochemical principles (Figure 21a). The total hydrogen production rate aligns within 2% of experimental data, further reinforcing the model’s predictive capability. Deviations may arise due to gas bubble formation effects, which introduce localized resistance and impact mass transport, partially accounted for in the model through porosity corrections in electrolyte conductivity. The model has been extended to include efficiency and energy consumption analysis (Figure 21b,c). Efficiency decreases as current density increases due to higher ohmic and activation losses. Peak efficiency occurs at low current densities but comes at the cost of lower hydrogen production rates. Efficiency is calculated as the ratio of the thermodynamic voltage (1.23 V) to the actual cell voltage. Energy consumption per mol of hydrogen increases with current density due to higher overpotentials. This highlights the trade-off between production rate and energy efficiency.
This Table 8 effectively summarizes the key comparisons and insights from the model versus experimental data, highlighting accuracy, deviations, and possible improvements. The comparisons above demonstrate that our FE model aligns well with experimental data, with deviations within acceptable ranges. These discrepancies can be attributed to factors such as simplifying assumptions in which the model assumes uniform temperature and neglects certain dynamic effects, which may not fully capture real-world complexities. Another is the parameter variability in which differences in membrane properties, catalyst loadings, and cell configurations between the model and experimental setups can lead to variations.
Our extended FEM model (Figure 22) effectively predicts PEM electrolyzer performance metrics, closely matching experimental work (EW) from reputable studies (EW(A) [272], EW(B) [271], and EW(C) [270]). The minor deviations observed are within acceptable limits and can be addressed in future work by incorporating more detailed physical phenomena and refining model parameters.
Although the absolute deviation values between the simulation and experimental results remain small, a slight increasing trend in deviation with rising current density is observed, particularly in Figure 22b. This behavior can be attributed to several factors: (i) nonlinearities become more prominent at higher current densities due to mass transport limitations and local heating effects that are only partially captured in the current model framework; (ii) membrane dehydration may occur under high load, affecting ohmic resistance, which the model presently treats as constant; (iii) gas accumulation effects near the electrode surface can locally increase overpotentials and are not explicitly resolved in the simulation; and (iv) minor numerical smoothing due to mesh resolution may underrepresent sharp gradients that develop at higher currents. These sources of deviation, while modest, offer insight into physical behaviors not fully captured by the current computational model and highlight areas for potential refinement in future simulations. At lower current densities, the deviations observed in Figure 22c,d can be attributed to underutilization of the electroactive area, slow activation of the electrode surface, and increased influence of membrane resistance and contact losses. These factors lead to divergence from the idealized model predictions. While such trends are well understood in electrochemical theory, direct experimental parallels for the specific datasets shown here are limited. Nevertheless, our extended FEM model aligns closely with experimental benchmarks [270,271], with average predictive error remaining within ±5% across the entire current density range.
In addition to conventional modeling techniques such as FEM, CFD, and system-level tools, recent studies have begun incorporating emerging computational approaches to address complex electrolysis challenges. Machine learning (ML) algorithms—including artificial neural networks, random forest models, and support vector regression—have shown promise in accelerating performance prediction, parameter tuning, and degradation forecasting. Meanwhile, dynamic control-oriented models, often implemented in platforms such as MATLAB/Simulink or Modelica, are being developed to simulate transient behavior, renewable coupling, and real-time control strategies. Multiscale modeling frameworks, which link electrochemical reaction kinetics with macroscopic transport phenomena, are also gaining traction for their ability to capture spatial and temporal variability across electrode and system domains. While these methods are still maturing, their integration into future electrolysis design and optimization workflows holds considerable potential.

6. Concluding Remarks

This review presents a comprehensive summary of advancements in electrolysis technologies for hydrogen production, focusing on material innovations, process optimizations, numerical modeling, and system integration. Electrolysis powered by renewable energy sources such as solar, wind, nuclear, and hydropower provides a sustainable pathway for large-scale hydrogen production. However, commercialization still faces challenges, including catalyst degradation, high energy consumption, operational stability, and economic constraints.
Significant progress has been made in electrode materials and electrocatalysts to enhance HER. Ni-based and modified Ni-alloy electrodes, as well as Mo, La, TiO2, Co, and metalloids combined with Ni, have improved catalytic efficiency and system longevity in AWE. In SOECs, perovskite composite electrodes such as CMF, Ni/YSZ, LSCM, SFFN/YSZ, and LSTO-SDC have demonstrated promising performance, particularly in co-electrolysis applications. Protective coatings, blocking layers, and reaction enhancements such as the reverse water gas shift (RWGS) reaction have further improved stability and conversion efficiencies. For PEM electrolysis, Pt and Ir catalysts remain dominant due to their high durability, with IrO2/Ti mitigating hydrogen degradation and F-doped IrO2 electrocatalysts enhancing electrochemical activity. Membrane optimization has also played a crucial role, where perfluorosulfonic acid ionomers outperform zirconium phosphate-modified membranes, and porous poly(perfluorosulfonic acid) membranes enhance ion conductivity over conventional membranes.
For MECs, advances in biocathode modifications, including PANI/MWCNT composites and Ni mesh cathodes, have improved biofilm formation and catalytic efficiency. Additionally, methane inhibition strategies and thermoelectric integration have shown potential in increasing MEC-driven hydrogen production efficiency. Alternative electrolytes, such as KH2PO4 for pressurized electrolysis, tetra-alkyl-ammonium-sulfonic acid for low temperature applications, and aqueous methanol for reduced operating voltages, provide new pathways for improving system efficiency and adaptability.
Beyond experimental advancements, numerical modeling and FEM-based simulations have proven to be powerful tools for understanding electrochemical kinetics, charge transport, and system optimization. PEM electrolysis, in particular, has benefited from multi-physics simulations, enabling predictive analyses of electrode behavior, current density effects, and efficiency improvements. These models provide valuable insights into performance optimization, guiding the design of next-generation electrolyzers.
Despite these advancements, future research should focus on pressure optimization, electrode durability, membrane improvements, and system integration with renewable energy sources to enhance hydrogen production efficiency and economic feasibility. Addressing scalability challenges, reducing operating temperatures, and refining reactor designs will be key to making electrolysis technologies commercially viable. Hybrid approaches, such as solar-assisted electrolysis, hybrid bio-electrochemical systems, and thermoelectric energy recovery, can further enhance efficiency and sustainability. With continued research and innovation, electrolysis has the potential to become a cornerstone technology for the global hydrogen economy, supporting large-scale storage for renewable energy derived electricity and long-term sustainability. Future research should focus on overcoming key technological bottlenecks such as catalyst degradation, membrane instability, and high system energy demand. Advanced materials for electrodes and membranes, pressure and temperature optimization strategies, and innovative cell designs will be critical to improving efficiency and durability. Additionally, integrating PEM and AWE systems with renewable energy sources, developing hybrid configurations, and leveraging digital tools such as AI-based modeling and real-time control systems offer promising pathways for scalable, flexible hydrogen production. These directions are vital to bridge the gap between laboratory performance and commercial deployment.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Midilli, A.; Kucuk, H.; Topal, M.E.; Akbulut, U.; Dincer, I. A comprehensive review on hydrogen production from coal gasification: Challenges and Opportunities. Int. J. Hydrogen Energy 2021, 46, 25385–25412. [Google Scholar] [CrossRef]
  2. Abe, J.O.; Popoola, A.P.I.; Ajenifuja, E.; Popoola, O.M. Hydrogen energy, economy and storage: Review and recommendation. Int. J. Hydrogen Energy 2019, 44, 15072–15086. [Google Scholar] [CrossRef]
  3. Dincer, I.; Siddiqui, O. 1.10 Heat Transfer Aspects of Energy. In Comprehensive Energy Systems; Dincer, I., Ed.; Elsevier: Oxford, UK, 2018; pp. 422–477. [Google Scholar] [CrossRef]
  4. David, M.; Ocampo-Martínez, C.; Sánchez-Peña, R. Advances in alkaline water electrolyzers: A review. J. Energy Storage 2019, 23, 392–403. [Google Scholar] [CrossRef]
  5. El-Shafie, M. Hydrogen production by water electrolysis technologies: A review. Results Eng. 2023, 20, 101426. [Google Scholar] [CrossRef]
  6. Temitayo, I.; Shree, O.B.; Akinwale, A.; Babalola, A.O. The integration of wind and solar power to water electrolyzer for green hydrogen production. Int. J. Hydrogen Energy 2024, 76, 75–96. [Google Scholar] [CrossRef]
  7. Smolinka, T.; Bergmann, H.; Garche, J.; Kusnezoff, M. The history of water electrolysis from its beginnings to the present. In Electrochemical Power Sources: Fundamentals, Systems, and Applications: Hydrogen Production by Water Electrolysis; Elsevier: Amsterdam, The Netherlands, 2022; pp. 83–164. [Google Scholar] [CrossRef]
  8. Temiz, M.; Kara, A.Y.G.; Erdemir, D.; Dincer, I. More efficient way of clean hydrogen production: The synergetic roles of magnetic effects and effective catalysts. Fuel 2024, 376, 132708. [Google Scholar] [CrossRef]
  9. Sharma, A.; Mehdi, S.E.H.; Pandit, S.; Eun-Oh, S.; Natarajan, V. Factors affecting hydrogen production in microbial electrolysis cell (MEC): A review. Int. J. Hydrogen Energy 2024, 61, 1473–1484. [Google Scholar] [CrossRef]
  10. Chakik, F.E.; Kaddami, M.; Mikou, M. Effect of operating parameters on hydrogen production by electrolysis of water. Int. J. Hydrogen Energy 2017, 42, 25550–25557. [Google Scholar] [CrossRef]
  11. May, M.; Basher, N. The Influence of Various Currents and Voltages on Hydrogen Production by Electrolysis Method. In Proceedings of the 7th International Conference on Science and Technology, Kyoto, Japan, 16 September 2024. [Google Scholar] [CrossRef]
  12. Jin, L.; Nakashima, R.N.; Comodi, C.; Frandsen, H.L. Alkaline electrolysis for green hydrogen production: A novel, simple model for thermo-electrochemical coupled system analysis. Appl. Therm. Eng. 2025, 262, 125154. [Google Scholar] [CrossRef]
  13. Kumar, S.S.; Himabindu, V. Hydrogen production by PEM water electrolysis—A review. Mater. Sci. Energy Technol. 2019, 2, 442–454. [Google Scholar] [CrossRef]
  14. Bhattacharyya, R.; Misra, A.; Sandeep, K.C. Photovoltaic solar energy conversion for hydrogen production by alkaline water electrolysis: Conceptual design and analysis. Energy Convers. Manag. 2017, 133, 1–13. [Google Scholar] [CrossRef]
  15. Bidin, N.; Azni, S.R.; Bakar, M.A.A.; Johari, A.R.; Munap, D.H.F.A.; Salebi, M.F.; Razak, S.N.A.; Sahidan, N.S.; Sulaiman, S.N.A. The effect of sunlight in hydrogen production from water electrolysis. Int. J. Hydrogen Energy 2017, 42, 133–142. [Google Scholar] [CrossRef]
  16. Giraldi, M.R.; François, J.-L.; Martin-del-Campo, C. Life cycle assessment of hydrogen production from a high temperature electrolysis process coupled to a high temperature gas nuclear reactor. Int. J. Hydrogen Energy 2015, 40, 4019–4033. [Google Scholar] [CrossRef]
  17. Bhandari, R.; Trudewind, C.A.; Zapp, P. Life cycle assessment of hydrogen production via electrolysis—A review. J. Clean. Prod. 2014, 85, 151–163. [Google Scholar] [CrossRef]
  18. Kwak, B.S.; Chae, J.; Kang, M. Design of a photochemical water electrolysis system based on a W-typed dye-sensitized serial solar module for high hydrogen production. Appl. Energy 2014, 125, 189–196. [Google Scholar] [CrossRef]
  19. Basile, A.; Iulianelli, A. (Eds.) Index. In Advances in Hydrogen Production, Storage and Distribution; Woodhead Publishing: Sawston, UK, 2014; pp. 525–546. [Google Scholar] [CrossRef]
  20. Wang, M.; Wang, Z.; Gong, X.; Guo, Z. The intensification technologies to water electrolysis for hydrogen production—A review. Renew. Sustain. Energy Rev. 2014, 29, 573–588. [Google Scholar] [CrossRef]
  21. Xie, W.-F.; Shao, M.-F. Alkaline Water Electrolysis for Efficient Hydrogen Production. J. Electrochem. 2022, 28, 2214008. [Google Scholar] [CrossRef]
  22. Tüysüz, H. Alkaline Water Electrolysis for Green Hydrogen Production. Acc. Chem. Res. 2024, 57, 558–567. [Google Scholar] [CrossRef]
  23. Hu, S.; Guo, B.; Ding, S.; Yang, F.; Dang, J.; Liu, B.; Gu, J.; Ma, J.; Ouyang, M. A comprehensive review of alkaline water electrolysis mathematical modeling. Appl. Energy 2022, 327, 120099. [Google Scholar] [CrossRef]
  24. Ehlers, J.C.; Feidenhans’l, A.A.; Therkildsen, K.T.; Larrazábal, G.O. Affordable Green Hydrogen from Alkaline Water Electrolysis: Key Research Needs from an Industrial Perspective. ACS Energy Lett. 2023, 8, 1502–1509. [Google Scholar] [CrossRef]
  25. Thissen, N.; Hoffmann, J.; Tigges, S.; Vogel, D.A.M.; Thoede, J.J.; Khan, S.; Schmitt, N.; Heumann, S.; Etzold, B.J.M.; Mechler, A.K. Industrially Relevant Conditions in Lab-Scale Analysis for Alkaline Water Electrolysis. ChemElectroChem 2024, 11, e202300432. [Google Scholar] [CrossRef]
  26. Dou, Z.; Rox, H.; Ramos, Z.; Baumann, R.; Ravishankar, R.; Czurratis, P.; Yang, X.; Lasagni, A.F.; Eckert, K.; Czarske, J.; et al. Scanning Acoustic Microscopy for Quantifying Two-phase Transfer in Operando Alkaline Water Electrolyzer. arXiv 2024, arXiv:2405.10716. [Google Scholar] [CrossRef]
  27. Brauns, J.; Turek, T. Alkaline Water Electrolysis Powered by Renewable Energy: A Review. Processes 2020, 8, 248. [Google Scholar] [CrossRef]
  28. Sebbahi, S.; Assila, A.; Belghiti, A.A.; Laasri, S.; Kaya, S.; Hlil, E.K.; Rachidi, S.; Hajjaji, A. A comprehensive review of recent advances in alkaline water electrolysis for hydrogen production. Int. J. Hydrogen Energy 2024, 82, 583–599. [Google Scholar] [CrossRef]
  29. Flis-Kabulska, I.; Sun, Y.; Zakroczymski, T.; Flis, J. Plasma carburizing for improvement of Ni-Fe cathodes for alkaline water electrolysis. Electrochim. Acta 2016, 220, 11–19. [Google Scholar] [CrossRef]
  30. Sivanantham, A.; Shanmugam, S. Nickel selenide supported on nickel foam as an efficient and durable non-precious electrocatalyst for the alkaline water electrolysis. Appl. Catal. B Environ. 2017, 203, 485–493. [Google Scholar] [CrossRef]
  31. Solmaz, R.; Döner, A.; Doğrubaş, M.; Erdoğan, İ.Y.; Kardaş, G. Enhancement of electrochemical activity of Raney-type NiZn coatings by modifying with PtRu binary deposits: Application for alkaline water electrolysis. Int. J. Hydrogen Energy 2016, 41, 1432–1440. [Google Scholar] [CrossRef]
  32. Buch, C.G.; Cardona, I.H.; Navarro, E.M.O.; Mestre, S.; Herranz, V.P. Synthesis and characterization of Au-modified macroporous Ni electrocatalysts for alkaline water electrolysis. Int. J. Hydrogen Energy 2016, 41, 764–772. [Google Scholar] [CrossRef]
  33. Rauscher, T.; Müller, C.I.; Schmidt, A.; Kieback, B.; Röntzsch, L. Ni–Mo–B alloys as cathode material for alkaline water electrolysis. Int. J. Hydrogen Energy 2016, 41, 2165–2176. [Google Scholar] [CrossRef]
  34. Kim, J.-H.; Lee, J.-N.; Yoo, C.-Y.; Lee, K.-B.; Lee, W.-M. Low-cost and energy-efficient asymmetric nickel electrode for alkaline water electrolysis. Int. J. Hydrogen Energy 2015, 40, 10720–10725. [Google Scholar] [CrossRef]
  35. Flis-Kabulska, I.; Flis, J.; Sun, Y.; Zakroczymski, T. Hydrogen evolution on plasma carburised nickel and effect of iron deposition from the electrolyte in alkaline water electrolysis. Electrochim. Acta 2015, 167, 61–68. [Google Scholar] [CrossRef]
  36. Cardoso, D.S.P.; Amaral, L.; Santos, D.M.F.; Šljukić, B.; Sequeira, C.A.C.; Macciò, D.; Saccone, A. Enhancement of hydrogen evolution in alkaline water electrolysis by using nickel-rare earth alloys. Int. J. Hydrogen Energy 2015, 40, 4295–4302. [Google Scholar] [CrossRef]
  37. Kjartansdóttir, C.; Caspersen, M.; Egelund, S.; Møller, P. Electrochemical investigation of surface area effects on PVD Al-Ni as electrocatalyst for alkaline water electrolysis. Electrochim. Acta 2014, 142, 324–335. [Google Scholar] [CrossRef]
  38. Tang, X.; Xiao, L.; Yang, C.; Lu, J.; Zhuang, L. Noble fabrication of Ni–Mo cathode for alkaline water electrolysis and alkaline polymer electrolyte water electrolysis. Int. J. Hydrogen Energy 2014, 39, 3055–3060. [Google Scholar] [CrossRef]
  39. Zuo, Y.; Bellani, S.; Ferri, M.; Saleh, G.; Shinde, D.V.; Zappia, M.I.; Brescia, R.; Prato, M.; De Trizio, L.; Infante, I.; et al. High-performance alkaline water electrolyzers based on Ru-perturbed Cu nanoplatelets cathode. Nat. Commun. 2023, 14, 4680. [Google Scholar] [CrossRef]
  40. Allebrod, F.; Chatzichristodoulou, C.; Mogensen, M. Cobalt and molybdenum activated electrodes in foam based alkaline electrolysis cells at 150–250 °C and 40 bar. J. Power Sources 2014, 255, 394–403. [Google Scholar] [CrossRef]
  41. Yüksel, H.; Özbay, A.; Solmaz, R.; Kahraman, M. Fabrication and characterization of three-dimensional silver nanodomes: Application for alkaline water electrolysis. Int. J. Hydrogen Energy 2017, 42, 2476–2484. [Google Scholar] [CrossRef]
  42. Liu, Y.; Wang, L.; Hübner, R.; Kresse, J.; Zhang, X.; Deconinick, M.; Vaynzof, Y.; Weidinger, I.M.; Eychmüller, A. Cobalt-based Co3Mo3N/Co4N/Co Metallic Heterostructure as a Highly Active Electrocatalyst for Alkaline Overall Water Splitting. Angew. Chem. Int. Ed. 2024, 63, e202319239. [Google Scholar] [CrossRef]
  43. Müller, C.I.; Rauscher, T.; Schmidt, A.; Schubert, T.; Weißgärber, T.; Kieback, B.; Röntzsch, L. Electrochemical investigations on amorphous Fe-base alloys for alkaline water electrolysis. Int. J. Hydrogen Energy 2014, 39, 8926–8937. [Google Scholar] [CrossRef]
  44. Lü, A.; Shen, Y.; Gong, X.; Wang, Z.; Wang, Y.; Wang, M. Effects of electrolyte recycling on desulfurization from bauxite water slurry electrolysis. Trans. Nonferrous Met. Soc. China 2016, 26, 1714–1720. [Google Scholar] [CrossRef]
  45. Yao, S.; Wang, C.; Li, M.; Cheng, Y.; Xing, W.; Jiang, S.P. Dimensionally stable Ni[sbnd]Fe@Co/Ti nanoporous electrodes by reactive deposition for water electrolysis. Int. J. Hydrogen Energy 2017, 42, 7143–7150. [Google Scholar] [CrossRef]
  46. Luo, S.; Zhang, T.; Xu, H.; Zhang, J.; Zhao, H.; Yun, J.; Zhao, H. Optimizing Alkaline Water Electrolysis: A Dual-Model Approach for Enhanced Hydrogen Production Efficiency. Energies 2024, 17, 5512. [Google Scholar] [CrossRef]
  47. Egert, F.; Ullmer, D.; Marx, S.; Taghizadeh, E.; Morawietz, T.; Gerle, M.; Le, T.A.; Schneider, L.P.C.; Biswas, I.S.; Wirz, R.E.; et al. Intensification of Alkaline Electrolyzer with Improved Two-Phase Flow. Adv. Energy Mater. 2025, 2405285. [Google Scholar] [CrossRef]
  48. Xia, Y.; Cheng, H.; He, H.; Wei, W. Efficiency and consistency enhancement for alkaline electrolyzers driven by renewable energy sources. Commun. Eng. 2023, 2, 22. [Google Scholar] [CrossRef]
  49. Nasser, M.; Megahed, T.F.; Ookawara, S.; Hassan, H. A review of water electrolysis–based systems for hydrogen production using hybrid/solar/wind energy systems. Environ. Sci. Pollut. Res. 2022, 29, 86994–87018. [Google Scholar] [CrossRef]
  50. Zeng, K.; Zhang, D. Evaluating the effect of surface modifications on Ni based electrodes for alkaline water electrolysis. Fuel 2014, 116, 692–698. [Google Scholar] [CrossRef]
  51. Ahn, S.H.; Park, H.-Y.; Choi, I.; Yoo, S.; Hwang, S.J.; Kim, H.-J.; Cho, E.; Yoon, C.; Park, H.; Son, H.; et al. Electrochemically fabricated NiCu alloy catalysts for hydrogen production in alkaline water electrolysis. Int. J. Hydrogen Energy 2013, 38, 13493–13501. [Google Scholar] [CrossRef]
  52. Hong, S.H.; Ahn, S.H.; Choi, J.; Kim, J.Y.; Kim, H.Y.; Kim, H.-J.; Jang, J.H.; Kim, H.; Kim, S.-K. High-activity electrodeposited NiW catalysts for hydrogen evolution in alkaline water electrolysis. Appl. Surf. Sci. 2015, 349, 629–635. [Google Scholar] [CrossRef]
  53. Egelund, S.; Caspersen, M.; Nikiforov, A.; Møller, P. Manufacturing of a LaNiO3 composite electrode for oxygen evolution in commercial alkaline water electrolysis. Int. J. Hydrogen Energy 2016, 41, 10152–10160. [Google Scholar] [CrossRef]
  54. Jović, B.M.; Lačnjevac, U.Č.; Jović, V.D.; Gajić-Krstajić, L.; Kovač, J.; Poleti, D.; Krstajić, N.V. Ni-(Ebonex-supported Ir) composite coatings as electrocatalysts for alkaline water electrolysis. Part II: Oxygen evolution. Int. J. Hydrogen Energy 2016, 41, 20502–20514. [Google Scholar] [CrossRef]
  55. He, H.; Chen, A.; Chang, M.; Ma, L.; Li, C. A feasible hydrogen evolution process of water electrolysis assisted by TiO2 nanotube photocatalysis. J. Ind. Eng. Chem. 2013, 19, 1112–1116. [Google Scholar] [CrossRef]
  56. Kuleshov, V.N.; Kuleshov, N.V.; Grigoriev, S.A.; Udris, E.Y.; Millet, P.; Grigoriev, A.S. Development and characterization of new nickel coatings for application in alkaline water electrolysis. Int. J. Hydrogen Energy 2016, 41, 36–45. [Google Scholar] [CrossRef]
  57. Li, Y.; Yang, S.; Li, H.; Li, G.; Li, M.; Shen, L.; Yang, Z.; Zhou, A. Electrodeposited ternary iron-cobalt-nickel catalyst on nickel foam for efficient water electrolysis at high current density. Colloids Surf. A Physicochem. Eng. Asp. 2016, 506, 694–702. [Google Scholar] [CrossRef]
  58. Li, Z.; Shi, S.; Zhong, Q.; Zhang, C.; Xu, C. Pt-Mn3O4/C as efficient electrocatalyst for oxygen evolution reaction in water electrolysis. Electrochim. Acta 2014, 146, 119–124. [Google Scholar] [CrossRef]
  59. Gao, R.; Zhang, H.; Yan, D. Iron diselenide nanoplatelets: Stable and efficient water-electrolysis catalysts. Nano Energy 2017, 31, 90–95. [Google Scholar] [CrossRef]
  60. Dobó, Z.; Palotas, A. Impact of the current fluctuation on the efficiency of Alkaline Water Electrolysis. Int. J. Hydrogen Energy 2016, 42, 5649–5656. [Google Scholar] [CrossRef]
  61. Berg, R.W.; Nikiforov, A.V.; Petrushina, I.M.; Bjerrum, N.J. Water vapor pressure over molten KH2PO4 and demonstration of water electrolysis at ∼300 °C. Appl. Energy 2016, 180, 269–275. [Google Scholar] [CrossRef]
  62. Tufa, R.A.; Rugiero, E.; Chanda, D.; Hnàt, J.; van Baak, W.; Veerman, J.; Fontananova, E.; Di Profio, G.; Drioli, E.; Bouzek, K.; et al. Salinity Gradient Power-Reverse Electrodialysis and Alkaline Polymer Electrolyte Water Electrolysis for hydrogen production. J. Membr. Sci. 2016, 514, 155–164. [Google Scholar] [CrossRef]
  63. Nikiforov, A.V.; Berg, R.W.; Petrushina, I.M.; Bjerrum, N.J. Specific electrical conductivity in molten potassium dihydrogen phosphate KH2PO4—An electrolyte for water electrolysis at ∼300 °C. Appl. Energy 2016, 175, 545–550. [Google Scholar] [CrossRef]
  64. Fiegenbaum, F.; Martini, E.; de Souza, M.O.; Becker, M.; Souza, R. Hydrogen production by water electrolysis using tetra-alkyl-ammonium-sulfonic acid ionic liquid electrolytes. J. Power Sources 2013, 243, 822–825. [Google Scholar] [CrossRef]
  65. Menia, S.; Tebibel, H.; Lassouane, F.; Khellaf, A.; Nouicer, I. Hydrogen production by methanol aqueous electrolysis using photovoltaic energy: Algerian potential. Int. J. Hydrogen Energy 2016, 42, 8661–8669. [Google Scholar] [CrossRef]
  66. Ge, L.; Gong, G.; Wang, Z.; Zhao, L.; Wang, Y. Insight of anode reaction for CWS (coal water slurry) electrolysis for hydrogen production. Energy 2016, 96, 372–382. [Google Scholar] [CrossRef]
  67. Chang, J.; Lv, Q.; Li, G.; Ge, J.; Liu, C.; Xing, W. Core-shell structured Ni12P5/Ni3(PO4)2 hollow spheres as difunctional and efficient electrocatalysts for overall water electrolysis. Appl. Catal. B Environ. 2017, 204, 486–496. [Google Scholar] [CrossRef]
  68. Li, L.; Laan, P.C.M.; Yan, X.; Cao, X.; Mekkering, M.J.; Zhao, K.; Ke, L.; Jiang, X.; Wu, X.; Li, L.; et al. High-Rate Alkaline Water Electrolysis at Industrially Relevant Conditions Enabled by Superaerophobic Electrode Assembly. Adv. Sci. 2023, 10, 2206180. [Google Scholar] [CrossRef]
  69. Jang, D.; Cho, H.-S.; Lee, S.; Park, M.; Kim, S.; Park, H.; Kang, S. Investigation of the operation characteristics and optimization of an alkaline water electrolysis system at high temperature and a high current density. J. Clean. Prod. 2023, 424, 138862. [Google Scholar] [CrossRef]
  70. Ding, S.; Guo, B.; Hu, S.; Gu, J.; Yang, F.; Li, Y.; Dang, J.; Liu, B.; Ma, J. Analysis of the effect of characteristic parameters and operating conditions on exergy efficiency of alkaline water electrolyzer. J. Power Sources 2022, 537, 231532. [Google Scholar] [CrossRef]
  71. Aili, D.; Hansen, M.K.; Andreasen, J.W.; Zhang, J.; Jensen, J.O.; Bjerrum, N.J.; Li, Q. Porous poly(perfluorosulfonic acid) membranes for alkaline water electrolysis. J. Membr. Sci. 2015, 493, 589–598. [Google Scholar] [CrossRef]
  72. Burnat, D.; Schlupp, M.; Wichser, A.; Lothenbach, B.; Gorbar, M.; Züttel, A.; Vogt, U.F. Composite membranes for alkaline electrolysis based on polysulfone and mineral fillers. J. Power Sources 2015, 291, 163–172. [Google Scholar] [CrossRef]
  73. Wu, Y.; Xu, G.; Zhou, J.; Cao, D. Research progress of the porous membranes in alkaline water electrolysis for green hydrogen production. Chem. Eng. J. 2025, 505, 159291. [Google Scholar] [CrossRef]
  74. Diaz, L.A.; Hnát, J.; Heredia, N.; Bruno, M.M.; Viva, F.A.; Paidar, M.; Corti, H.R.; Bouzek, K.; Abuin, G.C. Alkali doped poly (2,5-benzimidazole) membrane for alkaline water electrolysis: Characterization and performance. J. Power Sources 2016, 312, 128–136. [Google Scholar] [CrossRef]
  75. Suermann, M.; Schmidt, T.J.; Büchi, F.N. Cell Performance Determining Parameters in High Pressure Water Electrolysis. Electrochim. Acta 2016, 211, 989–997. [Google Scholar] [CrossRef]
  76. Koj, J.C.; Schreiber, A.; Zapp, P.; Marcuello, P. Life Cycle Assessment of Improved High Pressure Alkaline Electrolysis. Energy Procedia 2015, 75, 2871–2877. [Google Scholar] [CrossRef]
  77. Todd, D.; Schwager, M.; Mérida, W. Thermodynamics of high-temperature, high-pressure water electrolysis. J. Power Sources 2014, 269, 424–429. [Google Scholar] [CrossRef]
  78. Zouhri, K.; Lee, S. Evaluation and optimization of the alkaline water electrolysis ohmic polarization: Exergy study. Int. J. Hydrogen Energy 2016, 41, 7253–7263. [Google Scholar] [CrossRef]
  79. Milewski, J.; Guandalini, G.; Campanari, S. Modeling an alkaline electrolysis cell through reduced-order and loss-estimate approaches. J. Power Sources 2014, 269, 203–211. [Google Scholar] [CrossRef]
  80. Dahbi, S.; Aboutni, R.; Aziz, A.; Benazzi, N.; Elhafyani, M.; Kassmi, K. Optimised hydrogen production by a photovoltaic-electrolysis system DC/DC converter and water flow controller. Int. J. Hydrogen Energy 2016, 41, 20858–20866. [Google Scholar] [CrossRef]
  81. El-Askary, W.A.; Sakr, I.M.; Ibrahim, K.A.; Balabel, A. Hydrodynamics characteristics of hydrogen evolution process through electrolysis: Numerical and experimental studies. Energy 2015, 90, 722–737. [Google Scholar] [CrossRef]
  82. Tanaka, Y. Ion Exchange Membranes; Elsevier: Amsterdam, The Netherlands, 2015; Available online: https://shop.elsevier.com/books/ion-exchange-membranes/tanaka/978-0-444-63319-4 (accessed on 3 February 2025).
  83. Shestakova, M.; Bonete, P.; Gómez, R.; Sillanpää, M.; Tang, W.Z. Novel Ti/Ta2O5-SnO2 electrodes for water electrolysis and electrocatalytic oxidation of organics. Electrochim. Acta 2014, 120, 302–307. [Google Scholar] [CrossRef]
  84. Rakousky, C.; Reimer, U.; Wippermann, K.; Kuhri, S.; Carmo, M.; Lueke, W.; Stolten, D. Polymer electrolyte membrane water electrolysis: Restraining degradation in the presence of fluctuating power. J. Power Sources 2017, 342, 38–47. [Google Scholar] [CrossRef]
  85. Zhou, Y.; Zhong, H.; Chen, S.; Wen, G.; Shen, L.; Wang, Y.; Chen, R.; Tao, L.; Wang, S. Proton exchange membrane-based electrocatalytic systems for hydrogen production. Carbon Energy 2025, 7, e629. [Google Scholar] [CrossRef]
  86. Wang, C.R.; Stansberry, J.M.; Mukundan, R.; Chang, H.-M.J.; Kulkarni, D.; Park, A.M.; Plymill, A.B.; Firas, N.M.; Liu, C.P.; Lang, J.T.; et al. Proton Exchange Membrane (PEM) Water Electrolysis: Cell-Level Considerations for Gigawatt-Scale Deployment. Chem. Rev. 2025, 125, 1257–1302. [Google Scholar] [CrossRef] [PubMed]
  87. Fahr, S.; Engel, F.K.; Rehfeldt, S.; Peschel, A.; Klein, H. Overview and evaluation of crossover phenomena and mitigation measures in proton exchange membrane (PEM) electrolysis. Int. J. Hydrogen Energy 2024, 68, 705–721. [Google Scholar] [CrossRef]
  88. Rocha, F.; Georgiadis, C.; Van Droogenbroek, K.; Delmelle, R.; Pinon, X.; Pyka, G.; Kerckhofs, G.; Egert, F.; Razmjooei, F.; Ansar, S.-A.; et al. Proton exchange membrane-like alkaline water electrolysis using flow-engineered three-dimensional electrodes. Nat. Commun. 2024, 15, 7444. [Google Scholar] [CrossRef]
  89. Peng, X.; Deng, Z.; Zhao, X.; Li, G.; Song, J.; Liang, D.; Sun, X.; Xu, G.; Kang, W.; Liu, M. Experimental and Analytical Study of a Proton Exchange Membrane Electrolyser Integrated with Thermal Energy Storage for Performance Enhancement. Int. J. Photoenergy 2022, 2022, 7543121. [Google Scholar] [CrossRef]
  90. Araújo, H.F.; Gómez, J.A.; Santos, D.M.F. Proton-Exchange Membrane Electrolysis for Green Hydrogen Production: Fundamentals, Cost Breakdown, and Strategies to Minimize Platinum-Group Metal Content in Hydrogen Evolution Reaction Electrocatalysts. Catalysts 2024, 14, 845. [Google Scholar] [CrossRef]
  91. Wang, T.; Cao, X.; Jiao, L. PEM water electrolysis for hydrogen production: Fundamentals, advances, and prospects. Carbon Neutrality 2022, 1, 21. [Google Scholar] [CrossRef]
  92. Grigoriev, S.A.; Kalinnikov, A.A. Mathematical modeling and experimental study of the performance of PEM water electrolysis cell with different loadings of platinum metals in electrocatalytic layers. Int. J. Hydrogen Energy 2017, 42, 1590–1597. [Google Scholar] [CrossRef]
  93. Kúš, P.; Ostroverkh, A.; Ševčíková, K.; Khalakhan, I.; Fiala, R.; Skála, T.; Tsud, N.; Matolin, V. Magnetron sputtered Ir thin film on TiC-based support sublayer as low-loading anode catalyst for proton exchange membrane water electrolysis. Int. J. Hydrogen Energy 2016, 41, 15124–15132. [Google Scholar] [CrossRef]
  94. Rozain, C.; Mayousse, E.; Guillet, N.; Millet, P. Influence of iridium oxide loadings on the performance of PEM water electrolysis cells: Part I—Pure IrO2-based anodes. Appl. Catal. B Environ. 2016, 182, 153–160. [Google Scholar] [CrossRef]
  95. Rozain, C.; Mayousse, E.; Guillet, N.; Millet, P. Influence of iridium oxide loadings on the performance of PEM water electrolysis cells: Part II—Advanced oxygen electrodes. Appl. Catal. B Environ. 2016, 182, 123–131. [Google Scholar] [CrossRef]
  96. Lee, B.-S.; Ahn, S.H.; Park, H.-Y.; Choi, I.; Yoo, S.J.; Kim, H.-J.; Henkensmeier, D.; Kim, J.Y.; Park, S.; Nam, S.W.; et al. Development of electrodeposited IrO2 electrodes as anodes in polymer electrolyte membrane water electrolysis. Appl. Catal. B Environ. 2015, 179, 285–291. [Google Scholar] [CrossRef]
  97. Kadakia, K.; Datta, M.K.; Jampani, P.H.; Park, S.K.; Kumta, P.N. Novel F-doped IrO2 oxygen evolution electrocatalyst for PEM based water electrolysis. J. Power Sources 2013, 222, 313–317. [Google Scholar] [CrossRef]
  98. Lamy, C.; Jaubert, T.; Baranton, S.; Coutanceau, C. Clean hydrogen generation through the electrocatalytic oxidation of ethanol in a Proton Exchange Membrane Electrolysis Cell (PEMEC): Effect of the nature and structure of the catalytic anode. J. Power Sources 2014, 245, 927–936. [Google Scholar] [CrossRef]
  99. Kadakia, K.; Datta, M.K.; Velikokhatnyi, O.I.; Jampani, P.; Park, S.K.; Chung, S.J.; Kumta, P.N. High performance fluorine doped (Sn,Ru)O2 oxygen evolution reaction electro-catalysts for proton exchange membrane based water electrolysis. J. Power Sources 2014, 245, 362–370. [Google Scholar] [CrossRef]
  100. Sapountzi, F.M.; Gracia, J.M.; Weststrate, K.-J.; Fredriksson, H.O.A.; Niemantsverdriet, J.W. Electrocatalysts for the generation of hydrogen, oxygen and synthesis gas. Prog. Energy Combust. Sci. 2017, 58, 1–35. [Google Scholar] [CrossRef]
  101. Zlotorowicz, A.; Sunde, S.; Seland, F. Zirconium hydrogen phosphate as an additive in electrocatalytic layers for the oxygen evolution reaction in PEM water electrolysis. Int. J. Hydrogen Energy 2015, 40, 9982–9988. [Google Scholar] [CrossRef]
  102. Rakousky, C.; Reimer, U.; Wippermann, K.; Carmo, M.; Lueke, W.; Stolten, D. An analysis of degradation phenomena in polymer electrolyte membrane water electrolysis. J. Power Sources 2016, 326, 120–128. [Google Scholar] [CrossRef]
  103. Skulimowska, A.; Dupont, M.; Zaton, M.; Sunde, S.; Merlo, L.; Jones, D.J.; Rozière, J. Proton exchange membrane water electrolysis with short-side-chain Aquivion® membrane and IrO2 anode catalyst. Int. J. Hydrogen Energy 2014, 39, 6307–6316. [Google Scholar] [CrossRef]
  104. Wang, X.; Shao, Z.-G.; Li, G.; Zhang, L.; Zhao, Y.; Lu, W.; Yi, B. Preparation and characterization of partial-cocrystallized catalyst-coated membrane for solid polymer electrolyte water electrolysis. Int. J. Hydrogen Energy 2013, 38, 9057–9064. [Google Scholar] [CrossRef]
  105. Olesen, A.C.; Rømer, C.; Kær, S.K. A numerical study of the gas-liquid, two-phase flow maldistribution in the anode of a high pressure PEM water electrolysis cell. Int. J. Hydrogen Energy 2016, 41, 52–68. [Google Scholar] [CrossRef]
  106. Schalenbach, M.; Carmo, M.; Fritz, D.L.; Mergel, J.; Stolten, D. Pressurized PEM water electrolysis: Efficiency and gas crossover. Int. J. Hydrogen Energy 2013, 38, 14921–14933. [Google Scholar] [CrossRef]
  107. Lamy, C. From hydrogen production by water electrolysis to its utilization in a PEM fuel cell or in a SO fuel cell: Some considerations on the energy efficiencies. Int. J. Hydrogen Energy 2016, 41, 15415–15425. [Google Scholar] [CrossRef]
  108. Olivier, P.; Bourasseau, C.; Bouamama, B. Modelling, simulation and analysis of a PEM electrolysis system. IFAC-PapersOnLine 2016, 49, 1014–1019. [Google Scholar] [CrossRef]
  109. Langemann, M.; Fritz, D.L.; Müller, M.; Stolten, D. Validation and characterization of suitable materials for bipolar plates in PEM water electrolysis. Int. J. Hydrogen Energy 2015, 40, 11385–11391. [Google Scholar] [CrossRef]
  110. Millet, P. 9—Hydrogen production by polymer electrolyte membrane water electrolysis. In Compendium of Hydrogen Energy; Subramani, V., Basile, A., Veziroğlu, T.N., Eds.; Woodhead Publishing: Oxford, UK, 2015; pp. 255–286. [Google Scholar] [CrossRef]
  111. Hancke, R.; Bujlo, P.; Holm, T.; Ulleberg, Ø. High-pressure PEM water electrolyser performance up to 180 bar differential pressure. J. Power Sources 2024, 601, 234271. [Google Scholar] [CrossRef]
  112. Sun, S.; Shao, Z.; Yu, H.; Li, G.; Yi, B. Investigations on degradation of the long-term proton exchange membrane water electrolysis stack. J. Power Sources 2014, 267, 515–520. [Google Scholar] [CrossRef]
  113. Yilmaz, C.; Kanoglu, M. Thermodynamic evaluation of geothermal energy powered hydrogen production by PEM water electrolysis. Energy 2014, 69, 592–602. [Google Scholar] [CrossRef]
  114. Rozain, C.; Millet, P. Electrochemical characterization of Polymer Electrolyte Membrane Water Electrolysis Cells. Electrochim. Acta 2014, 131, 160–167. [Google Scholar] [CrossRef]
  115. Bensmann, B.; Hanke-Rauschenbach, R.; Sundmacher, K. In-situ measurement of hydrogen crossover in polymer electrolyte membrane water electrolysis. Int. J. Hydrogen Energy 2014, 39, 49–53. [Google Scholar] [CrossRef]
  116. Lim, C.K.; Liu, Q.; Zhou, J.; Sun, Q.; Chan, S.H. High-temperature electrolysis of synthetic seawater using solid oxide electrolyzer cells. J. Power Sources 2017, 342, 79–87. [Google Scholar] [CrossRef]
  117. Deka, D.J.; Gunduz, S.; Kim, J.; Fitzgerald, T.; Shi, Y.; Co, A.C.; Ozkan, U.S. Hydrogen Production from Water in a Solid Oxide Electrolysis Cell: Effect of Ni Doping on Lanthanum Strontium Ferrite Perovskite Cathodes. Ind. Eng. Chem. Res. 2019, 58, 22497–22505. [Google Scholar] [CrossRef]
  118. Mizusawa, T.; Araki, T.; Mori, M. Temperature and reactive current distributions in microtubular solid oxide electrolysis cells. Int. J. Hydrogen Energy 2016, 41, 13888–13900. [Google Scholar] [CrossRef]
  119. Hosoi, K.; Sakai, T.; Ida, S.; Ishihara, T. Ce0.6Mn0.3Fe0.1O2-δ as an Alternative Cathode Material for High Temperature Steam Electrolysis Using LaGaO3-based Oxide Electrolyte. Electrochim. Acta 2016, 194, 473–479. [Google Scholar] [CrossRef]
  120. Zhang, X.; O’Brien, J.E.; Tao, G.; Zhou, C.; Housley, G.K. Experimental design, operation, and results of a 4 kW high temperature steam electrolysis experiment. J. Power Sources 2015, 297, 90–97. [Google Scholar] [CrossRef]
  121. Reytier, M.; Di Iorio, S.; Chatroux, A.; Petitjean, M.; Cren, J.; De Saint Jean, M.; Aicart, J.; Mougin, J. Stack performances in high temperature steam electrolysis and co-electrolysis. Int. J. Hydrogen Energy 2015, 40, 11370–11377. [Google Scholar] [CrossRef]
  122. Alenazey, F.; Alyousef, Y.; Almisned, O.; Almutairi, G.; Ghouse, M.; Montinaro, D.; Ghigliazza, F. Production of synthesis gas (H2 and CO) by high-temperature Co-electrolysis of H2O and CO2. Int. J. Hydrogen Energy 2015, 40, 10274–10280. [Google Scholar] [CrossRef]
  123. Cacciuttolo, Q.; Vulliet, J.; Lair, V.; Cassir, M.; Ringuedé, A. Effect of pressure on high temperature steam electrolysis: Model and experimental tests. Int. J. Hydrogen Energy 2015, 40, 11378–11384. [Google Scholar] [CrossRef]
  124. Kazempoor, P.; Braun, R.J. Hydrogen and synthetic fuel production using high temperature solid oxide electrolysis cells (SOECs). Int. J. Hydrogen Energy 2015, 40, 3599–3612. [Google Scholar] [CrossRef]
  125. Ardigo, M.R.; Popa, I.; Combemale, L.; Chevalier, S.; Herbst, F.; Girardon, P. Dual atmosphere study of the K41X stainless steel for interconnect application in high temperature water vapour electrolysis. Int. J. Hydrogen Energy 2015, 40, 5305–5312. [Google Scholar] [CrossRef]
  126. Houaijia, A.; Breuer, S.; Thomey, D.; Brosig, C.; Säck, J.-P.; Roeb, M.; Sattler, C. Solar Hydrogen by High-temperature Electrolysis: Flowsheeting and Experimental Analysis of a Tube-type Receiver Concept for Superheated Steam Production. Energy Procedia 2014, 49, 1960–1969. [Google Scholar] [CrossRef]
  127. Mougin, J.; Chatroux, A.; Couturier, K.; Petitjean, M.; Reytier, M.; Gousseau, G.; Lefebvre-Joud, F. High Temperature Steam Electrolysis Stack with Enhanced Performance and Durability. Energy Procedia 2012, 29, 445–454. [Google Scholar] [CrossRef]
  128. Giddey, S.; Kulkarni, A.; Badwal, S.P.S. Low emission hydrogen generation through carbon assisted electrolysis. Int. J. Hydrogen Energy 2015, 40, 70–74. [Google Scholar] [CrossRef]
  129. Dong, D.; Shao, X.; Hu, X.; Chen, K.; Xie, K.; Yu, L.; Ye, Z.; Yang, P.; Parkinson, G.; Li, C.-Z. Improved gas diffusion within microchanneled cathode supports of SOECs for steam electrolysis. Int. J. Hydrogen Energy 2016, 41, 19829–19835. [Google Scholar] [CrossRef]
  130. Hauch, A.; Brodersen, K.; Chen, M.; Mogensen, M.B. Ni/YSZ electrodes structures optimized for increased electrolysis performance and durability. Solid State Ion. 2016, 293, 27–36. [Google Scholar] [CrossRef]
  131. Mahmood, A.; Bano, S.; Yu, J.H.; Lee, K.-H. Effect of operating conditions on the performance of solid electrolyte membrane reactor for steam and CO2 electrolysis. J. Membr. Sci. 2015, 473, 8–15. [Google Scholar] [CrossRef]
  132. Hou, X.; Zhao, K.; Yang, J.-I.; Ha, S. MoO2-based cathode for CO2 and H2O electrolysis. Int. J. Hydrogen Energy 2016, 41, 11895–11899. [Google Scholar] [CrossRef]
  133. Nechache, A.; Mansuy, A.; Petitjean, M.; Mougin, J.; Mauvy, F.; Boukamp, B.A.; Cassir, M.; Ringuedé, A. Diagnosis of a cathode-supported solid oxide electrolysis cell by electrochemical impedance spectroscopy. Electrochim. Acta 2016, 210, 596–605. [Google Scholar] [CrossRef]
  134. Li, H.; Chen, X.; Chen, S.; Wu, Y.; Xie, K. Composite manganate oxygen electrode enhanced with iron oxide nanocatalyst for high temperature steam electrolysis in a proton-conducting solid oxide electrolyzer. Int. J. Hydrogen Energy 2015, 40, 7920–7931. [Google Scholar] [CrossRef]
  135. Sirvent, J.D.D.; Carmona, A.; Rapenne, L.; Chiabrera, F.; Morata, A.; Burriel, M.; Baiutti, F.; Tarancón, A. Nanostructured La0.75Sr0.25Cr0.5Mn0.5O3–Ce0.8Sm0.2O2 Heterointerfaces as All-Ceramic Functional Layers for Solid Oxide Fuel Cell Applications. ACS Appl. Mater. Interfaces 2022, 14, 42178–42187. [Google Scholar] [CrossRef]
  136. Chen, S.; Xie, K.; Dong, D.; Li, H.; Qin, Q.; Zhang, Y.; Wu, Y. A composite cathode based on scandium-doped chromate for direct high-temperature steam electrolysis in a symmetric solid oxide electrolyzer. J. Power Sources 2015, 274, 718–729. [Google Scholar] [CrossRef]
  137. Li, H.; Sun, G.; Xie, K.; Qi, W.; Qin, Q.; Wei, H.; Chen, S.; Wang, Y.; Zhang, Y.; Wu, Y. Chromate cathode decorated with in-situ growth of copper nanocatalyst for high temperature carbon dioxide electrolysis. Int. J. Hydrogen Energy 2014, 39, 20888–20897. [Google Scholar] [CrossRef]
  138. Ge, B.; Ma, J.T.; Ai, D.; Deng, C.; Lin, X.; Xu, J. Sr2FeNbO6 Applied in Solid Oxide Electrolysis Cell as the Hydrogen Electrode: Kinetic Studies by Comparison with Ni-YSZ. Electrochim. Acta 2015, 151, 437–446. [Google Scholar] [CrossRef]
  139. Li, W.; Shi, Y.; Luo, Y.; Wang, Y.; Cai, N. Carbon deposition on patterned nickel/yttria stabilized zirconia electrodes for solid oxide fuel cell/solid oxide electrolysis cell modes. J. Power Sources 2015, 276, 26–31. [Google Scholar] [CrossRef]
  140. Keane, M.; Fan, H.; Han, M.; Singh, P. Role of initial microstructure on nickel-YSZ cathode degradation in solid oxide electrolysis cells. Int. J. Hydrogen Energy 2014, 39, 18718–18726. [Google Scholar] [CrossRef]
  141. Ruan, C.; Xie, K.; Yang, L.; Ding, B.; Wu, Y. Efficient carbon dioxide electrolysis in a symmetric solid oxide electrolyzer based on nanocatalyst-loaded chromate electrodes. Int. J. Hydrogen Energy 2014, 39, 10338–10348. [Google Scholar] [CrossRef]
  142. Xing, R.; Wang, Y.; Zhu, Y.; Liu, S.; Jin, C. Co-electrolysis of steam and CO2 in a solid oxide electrolysis cell with La0.75Sr0.25Cr0.5Mn0.5O3−δ–Cu ceramic composite electrode. J. Power Sources 2015, 274, 260–264. [Google Scholar] [CrossRef]
  143. Xu, S.; Dong, D.; Wang, Y.; Doherty, W.; Xie, K.; Wu, Y. Perovskite chromates cathode with resolved and anchored nickel nano-particles for direct high-temperature steam electrolysis. J. Power Sources 2014, 246, 346–355. [Google Scholar] [CrossRef]
  144. Xu, S.; Chen, S.; Li, M.; Xie, K.; Wang, Y.; Wu, Y. Composite cathode based on Fe-loaded LSCM for steam electrolysis in an oxide-ion-conducting solid oxide electrolyser. J. Power Sources 2013, 239, 332–340. [Google Scholar] [CrossRef]
  145. Dasari, H.P.; Park, S.-Y.; Kim, J.; Lee, J.-H.; Kim, B.-K.; Je, H.-J.; Lee, H.-W.; Yoon, K.J. Electrochemical characterization of Ni–yttria stabilized zirconia electrode for hydrogen production in solid oxide electrolysis cells. J. Power Sources 2013, 240, 721–728. [Google Scholar] [CrossRef]
  146. Li, Y.; Gan, Y.; Wang, Y.; Xie, K.; Wu, Y. Composite cathode based on Ni-loaded La0.75Sr0.25Cr0.5Mn0.5O3−δ for direct steam electrolysis in an oxide-ion-conducting solid oxide electrolyzer. Int. J. Hydrogen Energy 2013, 38, 10196–10207. [Google Scholar] [CrossRef]
  147. Yang, C.; Yang, Z.; Jin, C.; Liu, M.; Chen, F. High performance solid oxide electrolysis cells using Pr0.8Sr1.2(Co,Fe)0.8Nb0.2O4+δ–Co–Fe alloy hydrogen electrodes. Int. J. Hydrogen Energy 2013, 38, 11202–11208. [Google Scholar] [CrossRef]
  148. Liu, T.; Wang, Y.; Zhang, Y.; Fang, S.; Lei, L.; Ren, C.; Chen, F. Steam electrolysis in a solid oxide electrolysis cell fabricated by the phase-inversion tape casting method. Electrochem. Commun. 2015, 61, 106–109. [Google Scholar] [CrossRef]
  149. Li, Y.; Wu, G.; Ruan, C.; Zhou, Q.; Wang, Y.; Doherty, W.; Xie, K.; Wu, Y. Composite cathode based on doped vanadate enhanced with loaded metal nanoparticles for steam electrolysis. J. Power Sources 2014, 253, 349–359. [Google Scholar] [CrossRef]
  150. Gan, Y.; Qin, Q.; Chen, S.; Wang, Y.; Dong, D.; Xie, K.; Wu, Y. Composite cathode La0.4Sr0.4TiO3−δ–Ce0.8Sm0.2O2−δ impregnated with Ni for high-temperature steam electrolysis. J. Power Sources 2014, 245, 245–255. [Google Scholar] [CrossRef]
  151. Ardigo, M.R.; Popa, I.; Chevalier, S.; Parry, V.; Galerie, A.; Girardon, P.; Perry, F.; Laucournet, R.; Brevet, A.; Rigal, E. Coated interconnects development for high temperature water vapour electrolysis: Study in anode atmosphere. Int. J. Hydrogen Energy 2013, 38, 15910–15916. [Google Scholar] [CrossRef]
  152. Sumi, H.; Suda, E.; Mori, M. Blocking layer for prevention of current leakage for reversible solid oxide fuel cells and electrolysis cells with ceria-based electrolyte. Int. J. Hydrogen Energy 2017, 42, 4449–4455. [Google Scholar] [CrossRef]
  153. Pu, T.; Tan, W.; Shi, H.; Na, Y.; Lu, J.; Zhu, B. Steam/CO2 electrolysis in symmetric solid oxide electrolysis cell with barium cerate-carbonate composite electrolyte. Electrochim. Acta 2016, 190, 193–198. [Google Scholar] [CrossRef]
  154. Lauka, D.; Blumberga, D. Electrolysis Process Analysis by Using Low Carbon Content Additives: A Batch Test Study. Energy Procedia 2015, 72, 196–201. [Google Scholar] [CrossRef]
  155. Farandos, N.M.; Kleiminger, L.; Li, T.; Hankin, A.; Kelsall, G.H. Three-dimensional Inkjet Printed Solid Oxide Electrochemical Reactors. I. Yttria-stabilized Zirconia Electrolyte. Electrochim. Acta 2016, 213, 324–331. [Google Scholar] [CrossRef]
  156. Zhang, L.; Jie, X.; Shao, Z.-G.; Wang, X.; Yi, B. The dynamic-state effects of sodium ion contamination on the solid polymer electrolyte water electrolysis. J. Power Sources 2013, 241, 341–348. [Google Scholar] [CrossRef]
  157. Wang, X.; Shao, Z.-G.; Li, G.; Zhang, L.; Zhao, Y.; Lu, W.; Yi, B. A cocrystallized catalyst-coated membrane with high performance for solid polymer electrolyte water electrolysis. J. Power Sources 2013, 240, 525–529. [Google Scholar] [CrossRef]
  158. Xu, H.; Chen, B.; Irvine, J.; Ni, M. Modeling of CH4-assisted SOEC for H2O/CO2 co-electrolysis. Int. J. Hydrogen Energy 2016, 41, 21839–21849. [Google Scholar] [CrossRef]
  159. Tao, Y.; Ebbesen, S.D.; Mogensen, M.B. Degradation of solid oxide cells during co-electrolysis of steam and carbon dioxide at high current densities. J. Power Sources 2016, 328, 452–462. [Google Scholar] [CrossRef]
  160. Luo, Y.; Shi, Y.; Li, W.; Cai, N. Dynamic electro-thermal modeling of co-electrolysis of steam and carbon dioxide in a tubular solid oxide electrolysis cell. Energy 2015, 89, 637–647. [Google Scholar] [CrossRef]
  161. Menon, V.; Fu, Q.; Janardhanan, V.M.; Deutschmann, O. A model-based understanding of solid-oxide electrolysis cells (SOECs) for syngas production by H2O/CO2 co-electrolysis. J. Power Sources 2015, 274, 768–781. [Google Scholar] [CrossRef]
  162. Stempien, J.P.; Liu, Q.; Ni, M.; Sun, Q.; Chan, S.H. Physical principles for the calculation of equilibrium potential for co-electrolysis of steam and carbon dioxide in a Solid Oxide Electrolyzer Cell (SOEC). Electrochim. Acta 2014, 147, 490–497. [Google Scholar] [CrossRef]
  163. Luo, Y.; Shi, Y.; Li, W.; Cai, N. Comprehensive modeling of tubular solid oxide electrolysis cell for co-electrolysis of steam and carbon dioxide. Energy 2014, 70, 420–434. [Google Scholar] [CrossRef]
  164. Li, W.; Wang, H.; Shi, Y.; Cai, N. Performance and methane production characteristics of H2O–CO2 co-electrolysis in solid oxide electrolysis cells. Int. J. Hydrogen Energy 2013, 38, 11104–11109. [Google Scholar] [CrossRef]
  165. Kyriakou, V.; Garagounis, I.; Stoukides, M. Steam electrolysis with simultaneous production of C2 hydrocarbons in a solid electrolyte cell. Int. J. Hydrogen Energy 2014, 39, 675–683. [Google Scholar] [CrossRef]
  166. Patcharavorachot, Y.; Thongdee, S.; Saebea, D.; Authayanun, S.; Arpornwichanop, A. Performance comparison of solid oxide steam electrolysis cells with/without the addition of methane. Energy Convers. Manag. 2016, 120, 274–286. [Google Scholar] [CrossRef]
  167. Aicart, J.; Petitjean, M.; Laurencin, J.; Tallobre, L.; Dessemond, L. Accurate predictions of H2O and CO2 co-electrolysis outlet compositions in operation. Int. J. Hydrogen Energy 2015, 40, 3134–3148. [Google Scholar] [CrossRef]
  168. Li, W.; Shi, Y.; Luo, Y.; Cai, N. Elementary reaction modeling of CO2/H2O co-electrolysis cell considering effects of cathode thickness. J. Power Sources 2013, 243, 118–130. [Google Scholar] [CrossRef]
  169. Zheng, Y.; Xu, C.; Wang, W.G.; Guo, L. Quantitative electrochemical contributions of cells and stacked interfacial contacts in solid-oxide electrolysis cells. Int. J. Hydrogen Energy 2016, 41, 4538–4545. [Google Scholar] [CrossRef]
  170. Shimada, H.; Yamaguchi, T.; Kishimoto, H.; Sumi, H.; Yamaguchi, Y.; Nomura, K.; Fujishiro, Y. Nanocomposite electrodes for high current density over 3 A cm−2 in solid oxide electrolysis cells. Nat. Commun. 2019, 10, 5432. [Google Scholar] [CrossRef]
  171. Kim, S.; Joh, D.W.; Lee, D.-Y.; Lee, J.; Kim, H.S.; Khan, M.Z.; Hong, J.E.; Lee, S.-B.; Park, S.J.; Song, R.-H.; et al. Microstructure tailoring of solid oxide electrolysis cell air electrode to boost performance and long-term durability. Chem. Eng. J. 2021, 410, 128318. [Google Scholar] [CrossRef]
  172. Schefold, J.; Brisse, A.; Surrey, A.; Walter, C. 80,000 current on/off cycles in a one year long steam electrolysis test with a solid oxide cell. Int. J. Hydrogen Energy 2020, 45, 5143–5154. [Google Scholar] [CrossRef]
  173. Gaikwad, P.S.; Mondal, K.; Shin, Y.K.; van Duin, A.C.T.; Pawar, G. Enhancing the Faradaic efficiency of solid oxide electrolysis cells: Progress and perspective. npj Comput. Mater. 2023, 9, 149. [Google Scholar] [CrossRef]
  174. Jing, J.; Pang, J.; Chen, L.; Zhang, H.; Lei, Z.; Yang, Z. Structure,synthesis, properties and solid oxide electrolysis cells application of Ba(Ce, Zr)O3 based proton conducting materials. Chem. Eng. J. 2022, 429, 132314. [Google Scholar] [CrossRef]
  175. Laguna-Bercero, M.A.; Wang, Y.; Zhou, X.-D.; Zhu, L. Fundamentals of Solid Oxide Electrolysis Cells (SOEC). In High Temperature Electrolysis; Laguna-Bercero, M.A., Ed.; Springer International Publishing: Cham, Switzerland, 2023; pp. 5–34. [Google Scholar] [CrossRef]
  176. Xu, Y.; Cai, S.; Chi, B.; Tu, Z. Technological limitations and recent developments in a solid oxide electrolyzer cell: A review. Int. J. Hydrogen Energy 2024, 50, 548–591. [Google Scholar] [CrossRef]
  177. Wolf, S.E.; Winterhalder, F.E.; Vibhu, V.; de Haart, L.G.J.; Guillon, O.; Eichel, R.-A.; Menzler, N.H. Solid oxide electrolysis cells—Current material development and industrial application. J. Mater. Chem. A 2023, 11, 17977–18028. [Google Scholar] [CrossRef]
  178. Tanaka, Y.; Hoerlein, M.P.; Schiller, G. Numerical simulation of steam electrolysis with a solid oxide cell for proper evaluation of cell performances. Int. J. Hydrogen Energy 2016, 41, 752–763. [Google Scholar] [CrossRef]
  179. Sanz-Bermejo, J.; Muñoz-Antón, J.; Gonzalez-Aguilar, J.; Romero, M. Part load operation of a solid oxide electrolysis system for integration with renewable energy sources. Int. J. Hydrogen Energy 2015, 40, 8291–8303. [Google Scholar] [CrossRef]
  180. Peters, R.; Deja, R.; Blum, L.; Nguyen, V.N.; Fang, Q.; Stolten, D. Influence of operating parameters on overall system efficiencies using solid oxide electrolysis technology. Int. J. Hydrogen Energy 2015, 40, 7103–7113. [Google Scholar] [CrossRef]
  181. Zheng, Y.; Li, Q.; Chen, T.; Xu, C.; Wang, W.G. Quantitative contribution of resistance sources of components to stack performance for solid oxide electrolysis cells. J. Power Sources 2015, 274, 736–740. [Google Scholar] [CrossRef]
  182. Petrakopoulou, F.; Sanz-Bermejo, J.; Dufour, J.; Romero, M. Exergetic analysis of hybrid power plants with biomass and photovoltaics coupled with a solid-oxide electrolysis system. Energy 2016, 94, 304–315. [Google Scholar] [CrossRef]
  183. Kasai, S. Hydrogen electrical energy storage by high-temperature steam electrolysis for next-millennium energy security. Int. J. Hydrogen Energy 2014, 39, 21358–21370. [Google Scholar] [CrossRef]
  184. Cumming, D.J.; Elder, R.H. Thermal imaging of solid oxide cells operating under electrolysis conditions. J. Power Sources 2015, 280, 387–392. [Google Scholar] [CrossRef]
  185. Buttler, A.; Koltun, R.; Wolf, R.; Spliethoff, H. A detailed techno-economic analysis of heat integration in high temperature electrolysis for efficient hydrogen production. Int. J. Hydrogen Energy 2015, 40, 38–50. [Google Scholar] [CrossRef]
  186. Yu, X.; Shi, Y.; Wang, H.; Cai, N.; Li, C.; Tomov, R.I.; Hanna, J.; Glowacki, B.A.; Ghoniem, A.F. Experimental characterization and elementary reaction modeling of solid oxide electrolyte direct carbon fuel cell. J. Power Sources 2013, 243, 159–171. [Google Scholar] [CrossRef]
  187. Henke, M.; Willich, C.; Kallo, J.; Friedrich, K.A. Theoretical study on pressurized operation of solid oxide electrolysis cells. Int. J. Hydrogen Energy 2014, 39, 12434–12439. [Google Scholar] [CrossRef]
  188. Elder, R.; Cumming, D.; Mogensen, M.B. Chapter 11—High Temperature Electrolysis. In Carbon Dioxide Utilisation; Styring, P., Quadrelli, E.A., Armstrong, K., Eds.; Elsevier: Amsterdam, The Netherlands, 2015; pp. 183–209. [Google Scholar] [CrossRef]
  189. Mougin, J. 8—Hydrogen production by high-temperature steam electrolysis. In Compendium of Hydrogen Energy; Subramani, V., Basile, A., Veziroğlu, T.N., Eds.; Woodhead Publishing: Oxford, UK, 2015; pp. 225–253. [Google Scholar] [CrossRef]
  190. Nechache, A.; Cassir, M.; Ringuedé, A. Solid oxide electrolysis cell analysis by means of electrochemical impedance spectroscopy: A review. J. Power Sources 2014, 258, 164–181. [Google Scholar] [CrossRef]
  191. Wu, Y.; Shi, Y.; Luo, Y.; Cai, N. Elementary reaction modeling and experimental characterization of solid oxide direct carbon-assisted steam electrolysis cells. Solid State Ion. 2016, 295, 78–89. [Google Scholar] [CrossRef]
  192. Moçoteguy, P.; Brisse, A. A review and comprehensive analysis of degradation mechanisms of solid oxide electrolysis cells. Int. J. Hydrogen Energy 2013, 38, 15887–15902. [Google Scholar] [CrossRef]
  193. Jiang, Q.; Han, Y.; Tang, W.; Zhu, H.; Gao, C.; Chen, S.; Willander, M.; Cao, X.; Wang, Z.L. Self-powered seawater desalination and electrolysis using flowing kinetic energy. Nano Energy 2015, 15, 266–274. [Google Scholar] [CrossRef]
  194. Ji, X.; Liu, X.; Yang, W.; Xu, T.; Wang, X.; Zhang, X.; Wang, L.; Mao, X.; Wang, X. Sustainable phosphorus recovery from wastewater and fertilizer production in microbial electrolysis cells using the biochar-based cathode. Sci. Total Environ. 2022, 807, 150881. [Google Scholar] [CrossRef]
  195. N, S.; Spurgeon, J.; Matheswaran, M.; Satyavolu, J. Simultaneous biohydrogen production with distillery wastewater treatment using modified microbial electrolysis cell. Int. J. Hydrogen Energy 2020, 45, 18266–18274. [Google Scholar] [CrossRef]
  196. Seelajaroen, H.; Spiess, S.; Haberbauer, M.; Hassel, M.M.; Aljabour, A.; Thallner, S.; Guebitz, G.M.; Sariciftci, N.S. Enhanced methane producing microbial electrolysis cells for wastewater treatment using poly(neutral red) and chitosan modified electrodes. Sustain. Energy Fuels 2020, 4, 4238–4248. [Google Scholar] [CrossRef]
  197. Chung, T.H.; Meshref, M.N.A.; Hai, F.I.; Al-Mamun, A.; Dhar, B.R. Microbial electrochemical systems for hydrogen peroxide synthesis: Critical review of process optimization, prospective environmental applications, and challenges. Bioresour. Technol. 2020, 313, 123727. [Google Scholar] [CrossRef]
  198. Kadier, A.; Wang, J.; Chandrasekhar, K.; Abdeshahian, P.; Islam, M.A.; Ghanbari, F.; Bajpai, M.; Katoch, S.S.; Bhagawati, P.B.; Li, H.; et al. Performance optimization of microbial electrolysis cell (MEC) for palm oil mill effluent (POME) wastewater treatment and sustainable Bio-H2 production using response surface methodology (RSM). Int. J. Hydrogen Energy 2022, 47, 15464–15479. [Google Scholar] [CrossRef]
  199. Alcaraz-Gonzalez, V.; Flores-Estrella, R.A.; Nolasco, M.; Cano, V.; González-Alvarez, V. Microbial Electrolysis Cells for Biohydrogen Generation and Wastewater Treatment—A Short Review and Current Trends. In Wastewater Management and Technologies; Debik, E., Bahadir, M., Haarstrick, A., Eds.; Springer Nature: Cham, Switzerland, 2023; pp. 73–85. [Google Scholar] [CrossRef]
  200. Cui, W.; Yin, S. Microbial Electrolysis Cells for H2 Generation by Treating Acid Mine Drainage: Recent Advances and Emerging Trends. Fuels 2025, 6, 14. [Google Scholar] [CrossRef]
  201. Murugaiyan, J.; Narayanan, A.; Mohamed, S.N. An overview of microbial electrolysis cell configuration: Challenges and prospects on biohydrogen production. Int. J. Energy Res. 2022, 46, 20811–20827. [Google Scholar] [CrossRef]
  202. Koul, Y.; Devda, V.; Varjani, S.; Guo, W.; Ngo, H.H.; Taherzadeh, M.J.; Chang, J.-S.; Wong, J.W.C.; Bilal, M.; Kim, S.-H.; et al. Microbial electrolysis: A promising approach for treatment and resource recovery from industrial wastewater. Bioengineered 2022, 13, 8115–8134. [Google Scholar] [CrossRef] [PubMed]
  203. Radhika, D.; Shivakumar, A.; Kasai, D.R.; Koutavarapu, R.; Peera, S.G. Microbial Electrolysis Cell as a Diverse Technology: Overview of Prospective Applications, Advancements, and Challenges. Energies 2022, 15, 2611. [Google Scholar] [CrossRef]
  204. Cho, K.; Hoffmann, M.R. Molecular hydrogen production from wastewater electrolysis cell with multi-junction BiOx/TiO2 anode and stainless steel cathode: Current and energy efficiency. Appl. Catal. B Environ. 2017, 202, 671–682. [Google Scholar] [CrossRef]
  205. Lu, L.; Hou, D.; Fang, Y.; Huang, Y.; Ren, Z.J. Nickel based catalysts for highly efficient H2 evolution from wastewater in microbial electrolysis cells. Electrochim. Acta 2016, 206, 381–387. [Google Scholar] [CrossRef]
  206. Kadier, A.; Simayi, Y.; Chandrasekhar, K.; Ismail, M.; Kalil, M.S. Hydrogen gas production with an electroformed Ni mesh cathode catalysts in a single-chamber microbial electrolysis cell (MEC). Int. J. Hydrogen Energy 2015, 40, 14095–14103. [Google Scholar] [CrossRef]
  207. Lim, S.S.; Yu, E.H.; Daud, W.R.W.; Kim, B.H.; Scott, K. Bioanode as a limiting factor to biocathode performance in microbial electrolysis cells. Bioresour. Technol. 2017, 238, 313–324. [Google Scholar] [CrossRef]
  208. Chen, Y.; Xu, Y.; Chen, L.; Li, P.; Zhu, S.; Shen, S. Microbial electrolysis cells with polyaniline/multi-walled carbon nanotube-modified biocathodes. Energy 2015, 88, 377–384. [Google Scholar] [CrossRef]
  209. Li, J.-J.; Gao, M.-M.; Zhang, G.; Wang, X.-H.; Wang, S.-G.; Song, C.; Xu, Y.-Y. Perchlorate reduction in microbial electrolysis cell with polyaniline modified cathode. Bioresour. Technol. 2015, 177, 74–79. [Google Scholar] [CrossRef]
  210. Liu, J.; Zhang, F.; He, W.; Yang, W.; Feng, Y.; Logan, B.E. A microbial fluidized electrode electrolysis cell (MFEEC) for enhanced hydrogen production. J. Power Sources 2014, 271, 530–533. [Google Scholar] [CrossRef]
  211. Wang, Y.; Guo, W.-Q.; Xing, D.-F.; Chang, J.-S.; Ren, N.-Q. Hydrogen production using biocathode single-chamber microbial electrolysis cells fed by molasses wastewater at low temperature. Int. J. Hydrogen Energy 2014, 39, 19369–19375. [Google Scholar] [CrossRef]
  212. Huang, L.; Yao, B.; Wu, D.; Quan, X. Complete cobalt recovery from lithium cobalt oxide in self-driven microbial fuel cell—Microbial electrolysis cell systems. J. Power Sources 2014, 259, 54–64. [Google Scholar] [CrossRef]
  213. Cusick, R.D.; Ullery, M.L.; Dempsey, B.A.; Logan, B.E. Electrochemical struvite precipitation from digestate with a fluidized bed cathode microbial electrolysis cell. Water Res. 2014, 54, 297–306. [Google Scholar] [CrossRef]
  214. Marone, A.; Ayala-Campos, O.R.; Trably, E.; Carmona-Martínez, A.A.; Moscoviz, R.; Latrille, E.; Steyer, J.-P.; Alcaraz-Gonzalez, V.; Bernet, N. Coupling dark fermentation and microbial electrolysis to enhance bio-hydrogen production from agro-industrial wastewaters and by-products in a bio-refinery framework. Int. J. Hydrogen Energy 2017, 42, 1609–1621. [Google Scholar] [CrossRef]
  215. Huang, X.; Qu, Y.; Cid, C.A.; Finke, C.; Hoffmann, M.R.; Lim, K.; Jiang, S.C. Electrochemical disinfection of toilet wastewater using wastewater electrolysis cell. Water Res. 2016, 92, 164–172. [Google Scholar] [CrossRef]
  216. Dhar, B.R.; Elbeshbishy, E.; Hafez, H.; Lee, H.-S. Hydrogen production from sugar beet juice using an integrated biohydrogen process of dark fermentation and microbial electrolysis cell. Bioresour. Technol. 2015, 198, 223–230. [Google Scholar] [CrossRef]
  217. Heidrich, E.S.; Edwards, S.R.; Dolfing, J.; Cotterill, S.E.; Curtis, T.P. Performance of a pilot scale microbial electrolysis cell fed on domestic wastewater at ambient temperatures for a 12 month period. Bioresour. Technol. 2014, 173, 87–95. [Google Scholar] [CrossRef]
  218. Carmona-Martínez, A.A.; Trably, E.; Milferstedt, K.; Lacroix, R.; Etcheverry, L.; Bernet, N. Long-term continuous production of H2 in a microbial electrolysis cell (MEC) treating saline wastewater. Water Res. 2015, 81, 149–156. [Google Scholar] [CrossRef]
  219. Li, Y.; Styczynski, J.; Huang, Y.; Xu, Z.; McCutcheon, J.; Li, B. Energy-positive wastewater treatment and desalination in an integrated microbial desalination cell (MDC)-microbial electrolysis cell (MEC). J. Power Sources 2017, 356, 529–538. [Google Scholar] [CrossRef]
  220. Kuntke, P.; Sleutels, T.H.J.A.; Saakes, M.; Buisman, C.J.N. Hydrogen production and ammonium recovery from urine by a Microbial Electrolysis Cell. Int. J. Hydrogen Energy 2014, 39, 4771–4778. [Google Scholar] [CrossRef]
  221. Xu, L.; Liu, W.; Wu, Y.; Lee, P.; Wang, A.; Li, S. Trehalose enhancing microbial electrolysis cell for hydrogen generation in low temperature (0 °C). Bioresour. Technol. 2014, 166, 458–463. [Google Scholar] [CrossRef] [PubMed]
  222. Sun, R.; Xing, D.; Jia, J.; Liu, Q.; Zhou, A.; Bai, S.; Ren, N. Optimization of high-solid waste activated sludge concentration for hydrogen production in microbial electrolysis cells and microbial community diversity analysis. Int. J. Hydrogen Energy 2014, 39, 19912–19920. [Google Scholar] [CrossRef]
  223. Zhang, L.; Zhang, L.; Li, D. Enhanced dark fermentative hydrogen production by zero-valent iron activated carbon micro-electrolysis. Int. J. Hydrogen Energy 2015, 40, 12201–12208. [Google Scholar] [CrossRef]
  224. Wu, M.-S.; Ji, R.-Y.; Zheng, Y.-R. Nickel hydroxide electrode with a monolayer of nanocup arrays as an effective electrocatalyst for enhanced electrolysis of urea. Electrochim. Acta 2014, 144, 194–199. [Google Scholar] [CrossRef]
  225. Zhang, J.; Bai, Y.; Fan, Y.; Hou, H. Improved bio-hydrogen production from glucose by adding a specific methane inhibitor to microbial electrolysis cells with a double anode arrangement. J. Biosci. Bioeng. 2016, 122, 488–493. [Google Scholar] [CrossRef]
  226. Lewis, A.J.; Ren, S.; Ye, X.; Kim, P.; Labbe, N.; Borole, A.P. Hydrogen production from switchgrass via an integrated pyrolysis–microbial electrolysis process. Bioresour. Technol. 2015, 195, 231–241. [Google Scholar] [CrossRef]
  227. Watson, V.J.; Hatzell, M.; Logan, B.E. Hydrogen production from continuous flow, microbial reverse-electrodialysis electrolysis cells treating fermentation wastewater. Bioresour. Technol. 2015, 195, 51–56. [Google Scholar] [CrossRef]
  228. Song, Y.-H.; Hidayat, S.; Kim, H.-K.; Park, J.-Y. Hydrogen production in microbial reverse-electrodialysis electrolysis cells using a substrate without buffer solution. Bioresour. Technol. 2016, 210, 56–60. [Google Scholar] [CrossRef]
  229. Chen, Y.; Chen, M.; Shen, N.; Zeng, R.J. H2 production by the thermoelectric microconverter coupled with microbial electrolysis cell. Int. J. Hydrogen Energy 2016, 41, 22760–22768. [Google Scholar] [CrossRef]
  230. Montpart, N.; Baeza, M.; Baeza, J.A.; Guisasola, A. Low-cost fuel-cell based sensor of hydrogen production in lab scale microbial electrolysis cells. Int. J. Hydrogen Energy 2016, 41, 20465–20472. [Google Scholar] [CrossRef]
  231. Guo, K.; Prévoteau, A.; Rabaey, K. A novel tubular microbial electrolysis cell for high rate hydrogen production. J. Power Sources 2017, 356, 484–490. [Google Scholar] [CrossRef]
  232. Montpart, N.; Rago, L.; Baeza, J.A.; Guisasola, A. Hydrogen production in single chamber microbial electrolysis cells with different complex substrates. Water Res. 2015, 68, 601–615. [Google Scholar] [CrossRef] [PubMed]
  233. Luo, H.; Liu, G.; Zhang, R.; Bai, Y.; Fu, S.; Hou, Y. Heavy metal recovery combined with H2 production from artificial acid mine drainage using the microbial electrolysis cell. J. Hazard. Mater. 2014, 270, 153–159. [Google Scholar] [CrossRef] [PubMed]
  234. Zhang, J.; Zhang, Y.; Liu, B.; Dai, Y.; Quan, X.; Chen, S. A direct approach for enhancing the performance of a microbial electrolysis cell (MEC) combined anaerobic reactor by dosing ferric iron: Enrichment and isolation of Fe(III) reducing bacteria. Chem. Eng. J. 2014, 248, 223–229. [Google Scholar] [CrossRef]
  235. Chae, K.-J.; Kim, K.-Y.; Choi, M.-J.; Yang, E.; Kim, I.S.; Ren, X.; Lee, M. Sulfonated polyether ether ketone (SPEEK)-based composite proton exchange membrane reinforced with nanofibers for microbial electrolysis cells. Chem. Eng. J. 2014, 254, 393–398. [Google Scholar] [CrossRef]
  236. Khan, M.Z.; Nizami, A.S.; Rehan, M.; Ouda, O.K.M.; Sultana, S.; Ismail, I.M.; Shahzad, K. Microbial electrolysis cells for hydrogen production and urban wastewater treatment: A case study of Saudi Arabia. Appl. Energy 2017, 185, 410–420. [Google Scholar] [CrossRef]
  237. Cotterill, S.; Heidrich, E.; Curtis, T. 9—Microbial electrolysis cells for hydrogen production. In Microbial Electrochemical and Fuel Cells; Scott, K., Yu, E.H., Eds.; Woodhead Publishing: Boston, MA, USA, 2016; pp. 287–319. [Google Scholar] [CrossRef]
  238. Zhang, Y.; Angelidaki, I. Microbial electrolysis cells turning to be versatile technology: Recent advances and future challenges. Water Res. 2014, 56, 11–25. [Google Scholar] [CrossRef]
  239. Kadier, A.; Simayi, Y.; Kalil, M.S.; Abdeshahian, P.; Hamid, A.A. A review of the substrates used in microbial electrolysis cells (MECs) for producing sustainable and clean hydrogen gas. Renew. Energy 2014, 71, 466–472. [Google Scholar] [CrossRef]
  240. Rago, L.; Ruiz, Y.; Baeza, J.A.; Guisasola, A.; Cortés, P. Microbial community analysis in a long-term membrane-less microbial electrolysis cell with hydrogen and methane production. Bioelectrochemistry 2015, 106, 359–368. [Google Scholar] [CrossRef]
  241. Kadier, A.; Simayi, Y.; Abdeshahian, P.; Azman, N.F.; Chandrasekhar, K.; Kalil, M.S. A comprehensive review of microbial electrolysis cells (MEC) reactor designs and configurations for sustainable hydrogen gas production. Alex. Eng. J. 2016, 55, 427–443. [Google Scholar] [CrossRef]
  242. Kadier, A.; Kalil, M.S.; Abdeshahian, P.; Chandrasekhar, K.; Mohamed, A.; Azman, N.F.; Logroño, W.; Simayi, Y.; Hamid, A.A. Recent advances and emerging challenges in microbial electrolysis cells (MECs) for microbial production of hydrogen and value-added chemicals. Renew. Sustain. Energy Rev. 2016, 61, 501–525. [Google Scholar] [CrossRef]
  243. Chandran, P.; Bakshi, S.; Chatterjee, D. Study on the characteristics of hydrogen bubble formation and its transport during electrolysis of water. Chem. Eng. Sci. 2015, 138, 99–109. [Google Scholar] [CrossRef]
  244. Wang, Y.; Tie, J.; Tu, G.; Sun, S.; Zhao, R.; Zhang, Z. Effect of gas bubble on cell voltage oscillations based on equivalent circuit simulation in aluminum electrolysis cell. Trans. Nonferrous Met. Soc. China 2015, 25, 335–344. [Google Scholar] [CrossRef]
  245. Goupil, G.; Helle, S.; Davis, B.; Guay, D.; Roué, L. Anodic behavior of mechanically alloyed Cu–Ni–Fe and Cu–Ni–Fe–O electrodes for aluminum electrolysis in low-temperature KF-AlF3 electrolyte. Electrochim. Acta 2013, 112, 176–182. [Google Scholar] [CrossRef]
  246. Gutiérrez-Martín, F.; Ochoa-Mendoza, A.; Rodríguez-Antón, L.M. Pre-investigation of water electrolysis for flexible energy storage at large scales: The case of the Spanish power system. Int. J. Hydrogen Energy 2015, 40, 5544–5551. [Google Scholar] [CrossRef]
  247. Aghahosseini, S.; Dincer, I.; Naterer, G.F. Process integration of hydrolysis and electrolysis processes in the Cu–Cl cycle of hydrogen production. Int. J. Hydrogen Energy 2013, 38, 9633–9643. [Google Scholar] [CrossRef]
  248. Liu, S.; Zhou, W.; Tang, F.; Guo, B.; Zhang, Y.; Yin, R. Pretreatment of coal by ionic liquids towards coal electrolysis liquefaction. Fuel 2015, 160, 495–501. [Google Scholar] [CrossRef]
  249. Chen, Z.; Gong, X.; Wang, Z.; Wang, Y.; Zhang, S.; Xu, D. Sulfur removal from ionic liquid-assisted coal water slurry electrolysis in KNO3 system. J. Fuel Chem. Technol. 2013, 41, 928–936. [Google Scholar] [CrossRef]
  250. Šlampová, A.; Kubáň, P.; Boček, P. Additional considerations on electrolysis in electromembrane extraction. J. Chromatogr. A 2016, 1429, 364–368. [Google Scholar] [CrossRef]
  251. Cheng, Y.; Jiang, S.P. Advances in electrocatalysts for oxygen evolution reaction of water electrolysis-from metal oxides to carbon nanotubes. Prog. Nat. Sci. Mater. Int. 2015, 25, 545–553. [Google Scholar] [CrossRef]
  252. Guerra, L.; Gomes, J.; Puna, J.; Rodrigues, J. Preliminary study of synthesis gas production from water electrolysis, using the ELECTROFUEL® concept. Energy 2015, 89, 1050–1056. [Google Scholar] [CrossRef]
  253. Peng, S. (Ed.) Anion Exchange Membrane Water Electrolysis. In Electrochemical Hydrogen Production from Water Splitting: Basic, Materials and Progress; Springer Nature: Singapore, 2023; pp. 99–146. [Google Scholar] [CrossRef]
  254. Sugawara, Y.; Sankar, S.; Miyanishi, S.; Illathvalappil, R.; Gangadharan, P.K.; Kuroki, H.; Anilkumar, G.M.; Yamaguchi, T. Anion Exchange Membrane Water Electrolyzers: An Overview. J. Chem. Eng. Jpn. 2023, 56, 2210195. [Google Scholar] [CrossRef]
  255. Li, S.; Liu, T.; Zhang, W.; Wang, M.; Zhang, H.; Qin, C.; Zhang, L.; Chen, Y.; Jiang, S.; Liu, D.; et al. Highly efficient anion exchange membrane water electrolyzers via chromium-doped amorphous electrocatalysts. Nat. Commun. 2024, 15, 3416. [Google Scholar] [CrossRef] [PubMed]
  256. Lim, H.; Jeong, J.-Y.; Shin, G.; Kim, C.; Choi, G.; Myeong, S.-W.; Choi, S.M.; Park, T. Realizing Superior Durability of Water Electrolyzer Using Anion Exchange Membrane with an Interstitial Alkyl Chain: From a Single Cell to Large-Sized 1-cell Stack. Adv. Energy Mater. 2024, 14, 2401725. [Google Scholar] [CrossRef]
  257. Zheng, Y.; Ma, W.; Serban, A.; Allushi, A.; Hu, X. Anion Exchange Membrane Water Electrolysis at 10 A⋅cm−2 Over 800 Hours. Angew. Chem. Int. Ed. 2025, 64, e202413698. [Google Scholar] [CrossRef]
  258. Klingenhof, M.; Trzesniowski, H.; Koch, S.; Zhu, J.; Zeng, Z.; Metzler, L.; Klinger, A.; Elshamy, M.; Lehmann, F.; Buchheister, P.W.; et al. High-performance anion-exchange membrane water electrolysers using NiX (X = Fe,Co,Mn) catalyst-coated membranes with redox-active Ni–O ligands. Nat. Catal. 2024, 7, 1213–1222. [Google Scholar] [CrossRef]
  259. Xu, Q.; Zhang, L.; Zhang, J.; Wang, J.; Hu, Y.; Jiang, H.; Li, C. Anion Exchange Membrane Water Electrolyzer: Electrode Design, Lab-Scaled Testing System and Performance Evaluation. EnergyChem 2022, 4, 100087. [Google Scholar] [CrossRef]
  260. Chang, S.-C.; Gu, R.-E.; Chan, Y.-H. Parameter Analysis of Anion Exchange Membrane Water Electrolysis System by Numerical Simulation. Energies 2024, 17, 5682. [Google Scholar] [CrossRef]
  261. Motealleh, B.; Liu, Z.; Masel, R.I.; Sculley, J.P.; Ni, Z.R.; Meroueh, L. Next-generation anion exchange membrane water electrolyzers operating for commercially relevant lifetimes. Int. J. Hydrogen Energy 2021, 46, 3379–3386. [Google Scholar] [CrossRef]
  262. Liu, L.; Ma, H.; Khan, M.; Hsiao, B.S. Recent Advances and Challenges in Anion Exchange Membranes Development/Application for Water Electrolysis: A Review. Membranes 2024, 14, 85. [Google Scholar] [CrossRef]
  263. Yang, D.; Wang, H.; Li, Y.; Liu, M.; Han, J.; Huang, N. Simulation Research on the Performance of PEMEC with Single Channel Flow Field Established by Orthogonally Sinusoidal Function with Different Amplitudes and Angular Frequencies. Int. J. Energy Res. 2024, 2024, 5683908. [Google Scholar] [CrossRef]
  264. Tirumalasetti, P.R.; Weng, F.-B.; Dlamini, M.M.; Jung, G.-B.; Yu, J.-W.; Hung, C.C.; Nelli, D.; Hung, B.S.; Chiu, P.-C. A comparative numerical analysis of proton exchange membrane water electrolyzer using different flow field dynamics. Int. J. Hydrogen Energy 2024, 65, 572–581. [Google Scholar] [CrossRef]
  265. Kink, J.; Ise, M.; Bensmann, B.; Hanke-Rauschenbach, R. Modeling Mechanical Behavior of Membranes in Proton Exchange Membrane Water Electrolyzers. J. Electrochem. Soc. 2023, 170, 054507. [Google Scholar] [CrossRef]
  266. Ozdemir, S.N.; Taymaz, I.; San, F.G.B.; Okumuş, E. Performance assessment and optimization of the PEM water electrolyzer by coupled response surface methodology and finite element modeling. Fuel 2024, 365, 131138. [Google Scholar] [CrossRef]
  267. Hassan, A.H.; Wang, X.; Liao, Z.; Xu, C. Numerical Investigation on the Effects of Design Parameters and Operating Conditions on the Electrochemical Performance of Proton Exchange Membrane Water Electrolysis. J. Therm. Sci. 2023, 32, 1989–2007. [Google Scholar] [CrossRef]
  268. Jang, D.; Cho, H.-S.; Kang, S. Numerical modeling and analysis of the effect of pressure on the performance of an alkaline water electrolysis system. Appl. Energy 2021, 287, 116554. [Google Scholar] [CrossRef]
  269. Zheng, J.; Kang, Z.; Han, B.; Mo, J. Three-Dimensional Numerical Simulation of the Performance and Transport Phenomena of Oxygen Evolution Reactions in a Proton Exchange Membrane Water Electrolyzer. Materials 2023, 16, 1310. [Google Scholar] [CrossRef]
  270. Villagra, A.; Millet, P. An analysis of PEM water electrolysis cells operating at elevated current densities. Int. J. Hydrogen Energy 2019, 44, 9708–9717. [Google Scholar] [CrossRef]
  271. Schalenbach, M.; Tjarks, G.; Carmo, M.; Lueke, W.; Mueller, M.; Stolten, D. Acidic or Alkaline? Towards a New Perspective on the Efficiency of Water Electrolysis. J. Electrochem. Soc. 2016, 163, F3197. [Google Scholar] [CrossRef]
  272. Quemada-Villagómez, L.I.; Miranda-López, R.; Calderón-Ramírez, M.; Navarrete-Bolaños, J.L.; Martínez-González, G.M.; Jiménez-Islas, H. A simple and accurate mathematical model for estimating maximum and minimum daily environmental temperatures in a year. Build. Environ. 2021, 197, 107822. [Google Scholar] [CrossRef]
Figure 1. Classification of hydrogen production methods based on energy source and carbon footprint. The chart categorizes methods into thermochemical (e.g., steam methane reforming, coal gasification), electrochemical (e.g., water electrolysis using various power sources), and emerging technologies (e.g., photoelectrochemical and biological processes).
Figure 1. Classification of hydrogen production methods based on energy source and carbon footprint. The chart categorizes methods into thermochemical (e.g., steam methane reforming, coal gasification), electrochemical (e.g., water electrolysis using various power sources), and emerging technologies (e.g., photoelectrochemical and biological processes).
Eng 06 00081 g001
Figure 2. Workflow of the systematic literature review process used in this study.
Figure 2. Workflow of the systematic literature review process used in this study.
Eng 06 00081 g002
Figure 3. Classification of research papers based on (a) type of electrolysis, (b) anode material, (c) cathode material, and (d) electrolyte used.
Figure 3. Classification of research papers based on (a) type of electrolysis, (b) anode material, (c) cathode material, and (d) electrolyte used.
Eng 06 00081 g003
Figure 4. Classification of research papers based on type of (a) water membranes and (b) catalysts.
Figure 4. Classification of research papers based on type of (a) water membranes and (b) catalysts.
Eng 06 00081 g004
Figure 5. Classification of research papers based on (a) form of water and (b) temperature of the water.
Figure 5. Classification of research papers based on (a) form of water and (b) temperature of the water.
Eng 06 00081 g005
Figure 6. Schematic of the working principles of (a) AWE, (b) PEM electrolysis, (c) SOEC, and (d) MEC.
Figure 6. Schematic of the working principles of (a) AWE, (b) PEM electrolysis, (c) SOEC, and (d) MEC.
Eng 06 00081 g006
Figure 7. Mind map summarizing cathode optimization strategies in alkaline water electrolysis. The figure highlights material types, fabrication techniques, performance outcomes, trade-offs, and technology readiness levels.
Figure 7. Mind map summarizing cathode optimization strategies in alkaline water electrolysis. The figure highlights material types, fabrication techniques, performance outcomes, trade-offs, and technology readiness levels.
Eng 06 00081 g007
Figure 8. Effect of current density on energy efficiency in AWE. Efficiency trends reflect the thermodynamic energy conversion ratio based on HHV, showing performance decline at higher current densities due to increased overpotential. Data compiled from multiple studies conducted at ~80 °C in 25 wt.% KOH. (Data adapted from Refs. [60,61,62,66,68,69,70]).
Figure 8. Effect of current density on energy efficiency in AWE. Efficiency trends reflect the thermodynamic energy conversion ratio based on HHV, showing performance decline at higher current densities due to increased overpotential. Data compiled from multiple studies conducted at ~80 °C in 25 wt.% KOH. (Data adapted from Refs. [60,61,62,66,68,69,70]).
Eng 06 00081 g008
Figure 9. (a) Plot of the partial pressure of H2O with respect to temperature [61], (b) experimental setup [62] (c) plot between voltage and current in AWE [62], (d) plot between absolute temperature and conductivity for different molarities of electrolyte [63], and (e) plot between current density and the potential difference at different molarities of electrolyte [64]. (Adapted with permission).
Figure 9. (a) Plot of the partial pressure of H2O with respect to temperature [61], (b) experimental setup [62] (c) plot between voltage and current in AWE [62], (d) plot between absolute temperature and conductivity for different molarities of electrolyte [63], and (e) plot between current density and the potential difference at different molarities of electrolyte [64]. (Adapted with permission).
Eng 06 00081 g009
Figure 10. Effect of operating pressure on cell voltage in an alkaline water electrolyzer (AWE). The plot illustrates how increasing system pressure from 1 to 30 bar leads to a gradual rise in cell voltage due to increased gas solubility and partial pressure effects, which elevate the reversible potential and overpotentials.
Figure 10. Effect of operating pressure on cell voltage in an alkaline water electrolyzer (AWE). The plot illustrates how increasing system pressure from 1 to 30 bar leads to a gradual rise in cell voltage due to increased gas solubility and partial pressure effects, which elevate the reversible potential and overpotentials.
Eng 06 00081 g010
Figure 11. (a) Location of TiC support membrane in PEMWE [93], (b) curve between mass activity and anode loading [96], (c) plot between efficiency and time for different electrode compositions [95], (d) effect on efficiency at different IrO2 loadings [94], (e) curve between current density and cell voltage when the IrO2 membrane was doped with fluorine [97], and (f) curve between current density and cell voltage when RuO2 was used with IrO2 [99]. (Adapted with permission).
Figure 11. (a) Location of TiC support membrane in PEMWE [93], (b) curve between mass activity and anode loading [96], (c) plot between efficiency and time for different electrode compositions [95], (d) effect on efficiency at different IrO2 loadings [94], (e) curve between current density and cell voltage when the IrO2 membrane was doped with fluorine [97], and (f) curve between current density and cell voltage when RuO2 was used with IrO2 [99]. (Adapted with permission).
Eng 06 00081 g011aEng 06 00081 g011b
Figure 12. Performance characteristics of PEM water electrolyzers: (a) Effect of anode loading on mass activity, (b) Effect of operating pressure on PEMWE efficiency, (c) Influence of operating temperature on cell voltage, and (d) Long-term stability of PEMWE over time.
Figure 12. Performance characteristics of PEM water electrolyzers: (a) Effect of anode loading on mass activity, (b) Effect of operating pressure on PEMWE efficiency, (c) Influence of operating temperature on cell voltage, and (d) Long-term stability of PEMWE over time.
Eng 06 00081 g012
Figure 13. (a) Water electrolysis cell with the plot showing current efficiency and electrical efficiency [204], and (b) plot between potential and time at different cell operating conditions [207]. (Adapted with permission).
Figure 13. (a) Water electrolysis cell with the plot showing current efficiency and electrical efficiency [204], and (b) plot between potential and time at different cell operating conditions [207]. (Adapted with permission).
Eng 06 00081 g013
Figure 14. (a) Schematic representation of a two-stage biohydrogen production system integrating dark fermentation (Stage I) and a microbial electrolysis cell (Stage II) [214], (b) setup used for MEC [215], (c) integration of dark fermentation and MEC process for hydrogen production from sugar beet juice [216], (d) setup of a 100-L MEC operated for a 12-month period fed on raw domestic wastewater [217], (e) a biofilm-based 4 L two-chamber MEC continuously fed with acetate under saline conditions for more than 100 days [218], and (f) simultaneous removal of nitrogen in municipal wastewater [219]. (Adapted with permission).
Figure 14. (a) Schematic representation of a two-stage biohydrogen production system integrating dark fermentation (Stage I) and a microbial electrolysis cell (Stage II) [214], (b) setup used for MEC [215], (c) integration of dark fermentation and MEC process for hydrogen production from sugar beet juice [216], (d) setup of a 100-L MEC operated for a 12-month period fed on raw domestic wastewater [217], (e) a biofilm-based 4 L two-chamber MEC continuously fed with acetate under saline conditions for more than 100 days [218], and (f) simultaneous removal of nitrogen in municipal wastewater [219]. (Adapted with permission).
Eng 06 00081 g014aEng 06 00081 g014bEng 06 00081 g014c
Figure 15. Overview of modeling strategies applied to water electrolysis systems, highlighting typical application domains, strengths, and limitations of FEM, CFD, RSM, and system-level tools such as Aspen Plus®.
Figure 15. Overview of modeling strategies applied to water electrolysis systems, highlighting typical application domains, strengths, and limitations of FEM, CFD, RSM, and system-level tools such as Aspen Plus®.
Eng 06 00081 g015
Figure 16. Voltage gradient inside the electrolyte validating charge transport behavior in both 2D and 3D models. (2D map of electrochemical potential distribution and corresponding current density streamlines in a PEM electrolyzer simulation). The color gradient represents the potential field, while the white arrows show the direction and relative magnitude of current density vectors (J = −∇φ). Annotated regions with steep potential gradients illustrate where non-uniform current density distributions arise, providing visual support for the associated discussion in the text.
Figure 16. Voltage gradient inside the electrolyte validating charge transport behavior in both 2D and 3D models. (2D map of electrochemical potential distribution and corresponding current density streamlines in a PEM electrolyzer simulation). The color gradient represents the potential field, while the white arrows show the direction and relative magnitude of current density vectors (J = −∇φ). Annotated regions with steep potential gradients illustrate where non-uniform current density distributions arise, providing visual support for the associated discussion in the text.
Eng 06 00081 g016
Figure 17. Simulated overpotential contributions in a PEM electrolyzer model at 80 °C. The 2D plot shows the breakdown of activation (blue), ohmic (red), and concentration (green) overpotentials as a function of current density, with the total simulated cell voltage represented by a dashed black curve. The 3D surface visualization illustrates the spatial variation of activation and ohmic overpotentials across the cell domain. The red and blue planes correspond to ohmic and activation components, respectively. While the figure presents the simulation output, the total voltage curve generated by the model has been benchmarked against experimental polarization data from Villagra & Millet (2019) [270], showing close agreement across the full range of current densities. This supports the fidelity of the simulated outputs and confirms that the overpotential breakdown reflects realistic PEM cell behavior without requiring post-processed manual overlays.
Figure 17. Simulated overpotential contributions in a PEM electrolyzer model at 80 °C. The 2D plot shows the breakdown of activation (blue), ohmic (red), and concentration (green) overpotentials as a function of current density, with the total simulated cell voltage represented by a dashed black curve. The 3D surface visualization illustrates the spatial variation of activation and ohmic overpotentials across the cell domain. The red and blue planes correspond to ohmic and activation components, respectively. While the figure presents the simulation output, the total voltage curve generated by the model has been benchmarked against experimental polarization data from Villagra & Millet (2019) [270], showing close agreement across the full range of current densities. This supports the fidelity of the simulated outputs and confirms that the overpotential breakdown reflects realistic PEM cell behavior without requiring post-processed manual overlays.
Eng 06 00081 g017
Figure 18. Final gas fraction distribution at different time intervals.
Figure 18. Final gas fraction distribution at different time intervals.
Eng 06 00081 g018
Figure 19. 3D visualization of final gas volume fraction distribution in the electrolyzer domain. The color scale indicates local gas fraction intensity, with higher values near the central active region. The plot is normalized along the X and Y axes to represent the in-plane dimensions of the membrane electrode assembly (MEA). The edges of the domain (near X, Y ≈ 0, and 1) represent the electrode periphery, where gas accumulation tends to decrease due to radial dispersion and edge effects. The flow channels are aligned along the in-plane axes and promote gas evacuation across the domain. Gas concentration is highest near the center, indicating effective bubble retention in regions away from channel landings. This spatial mapping helps visualize mass transport gradients influenced by both geometry and flow-field configuration.
Figure 19. 3D visualization of final gas volume fraction distribution in the electrolyzer domain. The color scale indicates local gas fraction intensity, with higher values near the central active region. The plot is normalized along the X and Y axes to represent the in-plane dimensions of the membrane electrode assembly (MEA). The edges of the domain (near X, Y ≈ 0, and 1) represent the electrode periphery, where gas accumulation tends to decrease due to radial dispersion and edge effects. The flow channels are aligned along the in-plane axes and promote gas evacuation across the domain. Gas concentration is highest near the center, indicating effective bubble retention in regions away from channel landings. This spatial mapping helps visualize mass transport gradients influenced by both geometry and flow-field configuration.
Eng 06 00081 g019
Figure 20. (a) Polarization curve between cell voltage and current density and (b) overpotential contribution in PEM electrolysis.
Figure 20. (a) Polarization curve between cell voltage and current density and (b) overpotential contribution in PEM electrolysis.
Eng 06 00081 g020
Figure 21. (a) Analysis of hydrogen production rate, (b) electrolyzer efficiency, and (c) energy consumption with current density.
Figure 21. (a) Analysis of hydrogen production rate, (b) electrolyzer efficiency, and (c) energy consumption with current density.
Eng 06 00081 g021
Figure 22. Comparison between experimental results and model predictions: (a) Voltage gradient across the electrolyte, (b) Cell voltage versus current density, (c) Efficiency versus current density, and (d) Energy consumption versus current density for different electrolyte widths (EW(A), EW(B), and EW(C). The dashed blue lines represent model predictions.
Figure 22. Comparison between experimental results and model predictions: (a) Voltage gradient across the electrolyte, (b) Cell voltage versus current density, (c) Efficiency versus current density, and (d) Energy consumption versus current density for different electrolyte widths (EW(A), EW(B), and EW(C). The dashed blue lines represent model predictions.
Eng 06 00081 g022
Table 1. Methods of energy storage.
Table 1. Methods of energy storage.
S. No.Energy Storage MethodsPower RatingEfficiencyDischarge Time
1.High power supercapacitorup to 200 kW85–100%s–min
2.Superconducting magnetic storage500 kW–50 MW85–100%s–min
3.High power flywheelsup to 10 MW85–100%s–min
4.High energy supercapacitorup to 100 kW85–100%min–h
5.Long duration flywheelsup to 1 kW85–100%h–days
6.Batteries (including flow batteries)up to 50 MW70–85%h–days
7.Compressed air energy storage10 MW–1 GW45–70%h–days
8.Pumped hydro energy storage10 MW–1 GW70–85%days–month
9.Hydrogen and synthetic gas5 MW–1 GW30–45%days–month
10.Hydrogen with fuel cellsup to 1 MW30–45%days–month
Table 2. Variation in factors and their effect on electrolysis for hydrogen production.
Table 2. Variation in factors and their effect on electrolysis for hydrogen production.
S. No.FactorsVariationEffects on Electrolysis
1.Current densityIncreasedDirectly increases the rate of electrolysis.
2.TimeIncreasedInitially, activation phenomena (e.g., catalyst wetting, membrane conditioning) can cause a gradual increase in electrochemical performance. Over prolonged operation, performance degradation may occur, characterized by a decrease in current density at a given voltage, due to catalyst degradation, membrane aging, or electrolyte depletion or degradation.
3.Applied voltageIncreasedAn increase in driving force (applied potential) enhances the current density and accelerates the rate of electrolysis, up to a limit. Beyond this point, side reactions, mass transport limitations, and diffusion constraints may become significant, affecting efficiency and selectivity.
4.Concentration of ionsIncreasedGenerally, it increases conductivity and current density, enhancing the electrolysis rate. However, effects vary depending on the type of ions, membrane properties, and electrode materials involved.
5.Electrode active areaIncreasedIncreases total current capacity, potentially increasing the rate of electrolysis. However, current density (current per unit area) may decrease unless total current increases proportionally.
6.Distance between electrodesDecreasedReduces ohmic resistance, increasing current density and rate of electrolysis.
7.ResistanceReducedImproves conductivity, increasing current density and electrolysis rate.
8.TemperatureIncreasedEnhances ion mobility and reaction kinetics, increasing current density and rate of electrolysis.
9.PressureIncreasedReduces gas bubble size, lowering power consumption and increasing current density and electrolysis rate.
10.Electrode materialChanged from inert to activeActive electrodes may participate in reactions (e.g., oxidize), affecting stability and long-term efficiency.
11.Catalyst loading (mg/cm2)IncreasedCan enhance reaction kinetics and increase current density, improving electrolysis efficiency—up to an optimal level beyond which mass transport limitations may occur.
Table 3. The anode, cathode reactions, charge carrier, and overall cell reactions take place in AWE, PEM electrolysis, SOEC, and MEC.
Table 3. The anode, cathode reactions, charge carrier, and overall cell reactions take place in AWE, PEM electrolysis, SOEC, and MEC.
ElectrolysisAnode ReactionCathode ReactionCharge CarrierOverall Cell
AWE 2 O H 1 2 O 2 + H 2 O + 2 e H 2 O + 2 e H 2 + 2 O H O H H 2 O H 2 + 1 2 O 2
PEM H 2 O 2 H + + 1 2 O 2 + 2 e 2 H + + 2 e H 2 H + H 2 O H 2 + 1 2 O 2
SOEC O 2 1 2 O 2 + 2 e H 2 O + 2 e H 2 + O 2 O 2 H 2 O H 2 + 1 2 O 2
MEC C H 3 C O O + H 2 2 H C O 3 + 9 H + + 8 e 8 H + + 8 e 4 H 2 C H 3 C O O C H 3 O O + 4 H 2 O 2 H C O 3 + H + + 4 H 2
Table 4. Electrolyte materials used in SOEC.
Table 4. Electrolyte materials used in SOEC.
MaterialTemperature Range (°C)Ionic Conductivity (S/cm)Application
YSZ800–10000.01Conventional
ScSZ750–8500.015Improved performance
BCY500–6000.02Leakage prevention
GDC500–8000.012Low temperature
LSGM700–9000.018High performance
SPE800.005Water electrolysis
Table 5. Cathode materials used in SOEC.
Table 5. Cathode materials used in SOEC.
MaterialOperating Temperature (°C)Performance Improvement
Ni-YSZ800Stable but degrades
MoO2750Better for CO2 electrolysis
LSCM800High efficiency
LSV850Low resistance
SFN900Higher conductivity
LSTO800Catalytic improvement
Table 6. Some of the MEC’s electrode material with substrate, H2 production rate, and efficiency.
Table 6. Some of the MEC’s electrode material with substrate, H2 production rate, and efficiency.
Electrode MaterialSubstrate UsedH2 Production RateEfficiency
Stainless steel cathode and multi-junction semiconductor anode [204]Real wastewaterHigh EE (0.23),
CE (0.8)
Efficient at 200 A/m2
NiFe LDH electrocatalyst [205]WastewaterHigher H2 rateTwice as compared to stainless steel cathodes
Ni mesh cathode [206]Single chamber membrane-free MECHigh H2 productionAlternative to Pt/CC
Bioanode [207]Acetate7.4 L/dayStable at −1.0 V cathode potential
Biocathode [211]Low temperature wastewaterStable H2 generationOperated at 10 °C
Table 7. Assumptions and parameters for PEM electrolyzer.
Table 7. Assumptions and parameters for PEM electrolyzer.
ParameterValueUnits
Temperature60 °C°C
Exchange current density (i0)10−3 A/cm2A/cm2
Electrolyte conductivity (σ)0.1 S/cmS/cm
Cell resistance (R)0.2 Ω·cm2Ω·cm2
Transfer coefficient (α)0.5-
Number of electrons (n)2-
Faraday’s constant (F)96,485C/mol
Table 8. The key comparisons and insights from the model versus experimental data. (Experimental results refer to those reported in [270,271], which offer detailed electrochemical performance metrics for PEM water electrolyzers under conditions analogous to those used in our simulations. These datasets serve as the validation benchmark for evaluating the model’s predictive accuracy).
Table 8. The key comparisons and insights from the model versus experimental data. (Experimental results refer to those reported in [270,271], which offer detailed electrochemical performance metrics for PEM water electrolyzers under conditions analogous to those used in our simulations. These datasets serve as the validation benchmark for evaluating the model’s predictive accuracy).
Performance ParameterOur Model’s PredictionExperimental DataDeviation and Remarks
Cell voltage at 1 A/cm2~1.8 V~1.85 V [272]0.05 V (~2.7%) deviation, indicating good agreement.
Total overpotential at 1 A/cm2~0.57 V~0.6 V [271]0.03 V (~5%) deviation, suggesting reasonable accuracy.
Activation overpotentialFollows exponential trendShows similar trend in experiments~5% deviation due to approximations in exchange current density and reaction kinetics.
Ohmic overpotentialMatches experimental values closelyDeviates by <2%Reliable conductivity and resistance assumptions.
Concentration overpotentialMinimal but rises at higher current densitiesSimilar behavior was observed in experimentsThe model captures the trend effectively.
Hydrogen production rateIncreases linearly with current densityAligns within 2% of experimental dataSlight deviation due to gas bubble formation and localized resistance.
Efficiency at 1 A/cm2~68%~66% [272]2% deviation, demonstrating close alignment with experimental observations.
Efficiency trendDecreases with increasing current densityThe same trend was observed in experimentsConsistent with electrochemical principles.
Energy consumption vs. current densityIncreases due to higher overpotentialsExperimental results confirm trendThe trade-off between production rate and efficiency is captured well.
Discrepancies and limitationsDue to simplifying assumptions and parameter variabilityExperimental data affected by real-world complexitiesRefinements are needed to capture temperature variations, catalyst effects, and dynamic conditions.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vedrtnam, A.; Kalauni, K.; Pahwa, R. Water Electrolysis Technologies and Their Modeling Approaches: A Comprehensive Review. Eng 2025, 6, 81. https://doi.org/10.3390/eng6040081

AMA Style

Vedrtnam A, Kalauni K, Pahwa R. Water Electrolysis Technologies and Their Modeling Approaches: A Comprehensive Review. Eng. 2025; 6(4):81. https://doi.org/10.3390/eng6040081

Chicago/Turabian Style

Vedrtnam, Ajitanshu, Kishor Kalauni, and Rahul Pahwa. 2025. "Water Electrolysis Technologies and Their Modeling Approaches: A Comprehensive Review" Eng 6, no. 4: 81. https://doi.org/10.3390/eng6040081

APA Style

Vedrtnam, A., Kalauni, K., & Pahwa, R. (2025). Water Electrolysis Technologies and Their Modeling Approaches: A Comprehensive Review. Eng, 6(4), 81. https://doi.org/10.3390/eng6040081

Article Metrics

Back to TopTop