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 CO
2eq per kg H
2 [
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 H
2 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 CH
4 and H
2 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 Ni
3Se
2 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/cm
2 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 CH
4 + H
2 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/cm
2 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 Co
3Mo
3N/Co
4N/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 (Fe
82B
18, Fe
80Si
10B
10, and Fe
80Co
20Si
10B
10) 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/cm
2 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 LaNiO
3 electrode for OER in AWE. As a result, it was observed that the coating of LaNiO
3 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 TiO
2 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. NiCo
2O
4 spinels were deposited onto the anode to reduce the OER overpotential, while NiP
x 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-Mn
2O
4 catalysts with carbon black as a conductive agent and showed a remarkable activity for OER due to a synergistic effect between Mn
2O
4 and Pt. The Pt-Mn
2O
4/C catalyst with a weight ratio of Pt to Mn
2O
4 of 3:1 gives the best performance [
58]. Gao et al. synthesized 2D FeSe
2 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 (KH
2PO
4) as an electrolyte and demonstrated that at 300 °C, it dissociates into H
2O gas in equilibrium with KH
2PO
4, K
2H
2P
2O
7, and KPO
3, 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 KH
2PO
4 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 KH
2PO
4 at 300 °C. Hence, for temperatures as high as 300 °C, molten KH
2PO
4 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/cm
2), 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 CO
2 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 Ni
12P
5/Ni
3(PO
4)
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/cm
2, 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/cm
2 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/cm
2, 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 H
2O and temperature for KH
2PO
4. This was determined by Raman spectroscopy, using relative ratios of Raman scattering cross-sections of CH
4 and H
2O [
41]. As the temperature increases, the partial pressure of water also rises, showing an exponential trend, especially above 200 °C. The melting point of KH
2PO
4 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 Na
3PO
4, 3 M Na
3PO
4 + 2 M KH
2PO
4, 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/dm
3 and 4.2 mol/dm
3 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 H
2 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/Ta
2O
3-SnO
2 electrodes for H
2O 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/cm
2, while no significant degradation was observed at 1 A/cm
2, 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/cm
2), 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/cm
2, 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 H
2 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/cm
2 (or 1 mg/cm
2 of Pt/C with 40 wt.% Pt) and an IrO
2 loading of approximately 2.5 mg/cm
2 [
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/cm
3 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 IrO
2 particles at the anode of PEMWE cells. According to MEA, for over 1000 h of operation, IrO
2/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 IrO
2 anode) to only 20 µV h
−1 (for the 50 wt.% IrO
2/Ti anode) [
94]. Rozain et al. have studied PEMWE using IrO
2 alone as a catalyst at 80 °C. They used 0.25 mg/cm
2 of Pt cathode and 0.5 mg/cm
2 of IrO
2 anode at 1 A/cm
2 of current density with a Nafion 115 membrane [
95]. Lee et al. have also developed electrodeposited IrO
2 anodes in PEMWE. At 0.7 V, the IrO
2 loading ranges from 0.007 to 0.464 mg/cm
2 at 90 °C in the PEMWE test. An IrO
2 loading of 0.1 mg/cm
2 with the IrO
2/CP electrode shows the highest performance [
96]. Kadakia et al. investigated a nanostructured fluorine (F)-doped IrO
2 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 IrO
2 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/cm
2 for a 220 cm
3 evolution rate of H
2 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 SnO
2 with 10 wt.% or 20 wt.% RuO
2 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/cm
2 of pure IrO
2 and 0.12 mg/cm
2 of IrO
2/Ti particles. It is clear that IrO
2/Ti shows better efficiency as compared to pure IrO
2.
Figure 11d shows the curve of how efficiency is affected by different IrO
2 loadings. It is clear from the plot that 0.5 mg/cm
2 of IrO
2 shows the maximum efficiency of 71%.
Figure 11e displays the curve between the cell voltage and the current density when the IrO
2 membrane was doped with fluorine, and
Figure 11f demonstrates the curve between the cell voltage and the current density when RuO
2 was used with IrO
2.
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/cm
2 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 H
2 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 H
2 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 H
2 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 H
2 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 O
2− 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 H
2 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 (La
0.7Sr
0.2FeO
3) was evaluated as an SOEC cathode material. Electrolysis was performed at 800 °C with a 3% H
2O/He gas stream across various current densities. The study revealed that the undoped La
0.7Sr
0.2FeO
3 exhibited reduced Faradaic efficiency for hydrogen production due to the formation of a non-conductive La
2O
3 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 Ce
0.6Mn
0.3Fe
0.1O
2− (CMF) as a new oxide cathode in SOEC using LaGaO
3-based electrolyte for H
2 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/cm
2 [
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 Nm
3/h of H
2 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 H
2 by SOEC and observed that the production of H
2O and CO
2 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 H
2 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(H
2O) and high current densities. Applying such conditions, they showed a 0.3–0.4%/kh degradation rate of 1 A/cm
2 at 800 °C [
130]. Mahmood et al. developed a YSZ electrolyte membrane for a solid electrolyte membrane reactor and used electrochemical activity of CO
2, steam, and a CO
2-steam mixture at 800 °C. A very high current density was detected for CO
2, steam, and CO
2-steam mixture electrolysis at 850 °C and 1.5 V [
131]. Hou et al. used an MoO
2-based cathode in their experiment of co-electrolysis based on SOEC at 750 °C. It was observed that, the MoO
2-based and the Ni/YSZ cells showed alike electrochemical performance in the co-electrolysis mode, while in the CO
2 electrolysis mode, the MoO
2 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 La
0.8Sr
0.2MnO
3− (LSM) in SOEC and reported the maximum current efficiency of 65% of Fe
2O
3-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 CO
2 electrolysis in an oxide-ion-conducting SOE [
137]. Ge et al. introduced Sr
2FeNbO
6 (SFN) to the SOEC as the H
2 electrode. Higher conductivity was observed in the H
2/H
2O atmosphere as compared to that in the air. Also, if compared with Ni/YSZ, SFN-YSZ was more appropriate for the H
2 electrode in SOEC. Ge et al. introduced Sr
2FeNbO
6 (SFN) as a hydrogen electrode material in SOEC, reporting high electronic conductivity (2.215 S·cm
−1 at 850 °C in H
2/H
2O), 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 CO
2 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 CO
2 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 H
2 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 K
2NiF
4-type structured Pr
0.8Sr
1.2(CoFe)
0.8Nb
0.2O
4+ (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 H
2 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 H
2 [
149]. Gan et al. observed that the electro-catalytic activity of La
0.4Sr
0.4TiO
3 (LSTO) was insufficient for efficient electrochemical reduction of steam or CO
2. 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 La
0.8Sr
0.2MnO
2− and LaNi
0.6Fe
0.4O
2− as coatings in an O
2-H
2O atmosphere at 800 °C. The results concluded that LaNi
0.6Fe
0.4O
2− 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 BaCe
0.8Y
0.2O
3− (BCY) between Ce
0.9Gd
0.1O
1.95 (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 O
2 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 BaCe
0.8Y
0.2O
3− 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/cm
2 converting 31% of a combination of H
2:H
2O:CO
2 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 H
2O and CO
2 co-electrolysis in an SOEC is capable of energy storage and utilization of CO
2. Xu et al. numerically studied the effects of CH
4 on CO
2/H
2O co-electrolysis by a 2D model. It was observed that the CH
4 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/cm
2. 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 CO
2/H
2O 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 H
2O and CO
2 SOEC co-electrolysis. It was concluded that at temperatures above 800 °C, when the concentration of CO
2 is below 25%, then the oxygen partial pressure model suffices for modern cells [
162]. Luo et al. analyzed the efficiency and the performance of H
2O/CO
2 co-electrolysis in TSOEC. The results showed that the CO
2 conversion ratio was significantly promoted by the reversed water gas shift reaction [
163]. Li et al. have tested the H
2O-CO
2 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 CH
4 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 (O
2−) 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 CH
4. 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 H
2O/CO
2 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, H
2 electrode contact, and the electrochemical contributions of the cell at 750 °C in the SOEC stack with a 90/10 H
2O/H
2 ratio. At a constant current density, the air electrode contact, H
2 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/cm
2 is feasible using nanocomposite oxygen electrodes fabricated via spray pyrolysis. The bimodal-structured electrodes composed of Sm
0.5Sr
0.5CoO
3−δ and Ce
0.8Sm
0.2O
1.9 enabled high performance due to improved conductivity and gas diffusion. These cells achieved 3.13 A/cm
2 at 750 °C and 4.08 A/cm
2 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/cm
2 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 O
2−- 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/cm
2 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 H
2 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 H
2 by SOEC. The calculated efficiencies vary from 104% to 62% on the basis of lower heating value of H
2 produced [
180]. Zheng et al. investigated the NiO/YSZ/GDC/LSCF-GDC SOEC stack at the H
2O/H
2 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 H
2 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 H
2 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 H
2-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 CO
2 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/m
2 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 Sr
2FeNbO
6 (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/cm
2, 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 MoO
2, 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 H
2O and CO
2 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 (H
2) 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 CO
2. 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 Fe
2+-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/m
2. 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 CH
4/g
COD), 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 H
2O
2 synthesis via the cathodic oxygen reduction reaction. Their review emphasized the environmental applications of bio-electrochemically generated H
2O
2, 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/m
2 of real wastewater [
204]. Lu et al. produced a NiFe layered double hydroxide (NiFe LDH) electrocatalyst on NF for H
2 production from wastewater. The new cathode showed a high H
2 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 H
2 production in a single chamber membrane-free MEC [
206]. Lim et al. stated the bioanode at a current density of 0.36 A/m
2 and 0.37 A/m
2 with the potential of 0 and +0.6 V, respectively. During this, the H
2 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 H
2 production in single chamber, membrane-free biocathode MECs. The results concluded that with an increment in applied voltage, the H
2 production rates increased, and the performance of MECs was improved with modified biocathodes at a higher current density and H
2 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 H
2 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 H
2 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 LiCoO
2 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-H
2 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 H
2 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 H
2 [
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 H
2 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 H
2 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 H
2 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 H
2 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 H
2 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% H
2 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 H
2 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 H
2 production from glucose by adding CH
4. It was observed that the chloroform was an effective methane inhibitor and improved the efficiency of H
2 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 H
2 from biomass [
226]. Watson et al. used a microbial reverse-electrodialysis electrolysis cell (MREC) from the fermentation of wastewater to yield H
2. The results confirmed that if effluent anolyte COD concentrations were sufficient for anode potential, then the consistent rates of H
2 produced by MREC [
227]. Song et al. have used substrate (without buffer solution) under continuous flow conditions in an MREC for H
2 production. The hydrogen produced was 0.61 m
3-H
2, 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 H
2 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 H
2 [
229]. As a monitoring tool for hydrogen-producing MEC, Montpart et al. presented a low-cost sensor for H
2 production measurement. As a result, the fuel cell electrical signal had a high correlation with H
2 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 H
2 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 H
2 recovery, high H
2 purity, and outstanding operational stability [
231]. Montpart et al. have also used a single chamber MEC for H
2 production from synthetic wastewater. The results displayed that at 0.8 V, the current intensity, H
2 production, and cathodic gas recovery were 150 A/m
2, 0.94 H
2m
3/m
3d, and 91%, respectively [
232]. Lou et al. utilized a dual-chamber MEC for H
2 production, and also metal removal, such as Cu
2+, Ni
2+, and Fe
2+ 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 H
2 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 H
2 production at the cathode, with overall H
2 efficiency as compared to Nafion [
235].
Khan et al. reviewed the status of MEC as a mean wastewater treatment method and H
2 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 H
2 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 H
2 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 H
2 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 H
2 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 H
2 and current production rate in MECs, because the membrane-free design can lead to both on H
2 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.